{ "0801/0801.2870_arXiv.txt": { "abstract": "{The SKA will be a unique instrument with which to study the evolution of the gas content of galaxies. A proposed deep ($\\sim$ 8 Msec) `pencil-beam' survey is simulated using recently updated specifications for SKA sensitivity and survey speed. Almost $10^7$ galaxies could be detected in the redshifted 21cm line, most at redshifts in excess of two. This will enable confident statements to be made about the evolution of the cosmic HI density and the HI mass function to $z=3$, corresponding to a lookback time of 11 Gyr. However, galaxies or groups of galaxies with masses the same as the most HI-massive galaxies at $z=0$ will be detectable at redshifts of 6, if they exist. The ideal instrument for studying HI evolution would have an instantaneous sensitivity at least a factor of two higher than current specifications in the critical frequency range 200-500 MHz, or $A/T>2\\times 10^4$ m$^2$ K$^{-1}$. The capabilities of the SKA will be highly complementary to ALMA which will be able to study the evolution of the molecular gas component over the same redshift range.} \\FullConference{From planets to dark energy: the modern radio universe\\\\ October 1-5 2007\\\\ University of Manchester, Manchester, UK} \\begin{document} ", "introduction": "Understanding the evolution of galaxies is one of the goals of modern cosmology and one of the five key science goals of the SKA \\cite{cr04}. A key ingredient in galaxy evolution is the gas supplied through various accretion, merger and feedback processes that occur during the assembly of galaxies. This gas, which is mainly hydrogen, passes through a neutral atomic phase and later condenses into massive molecular clouds and stars. During the atomic phase, it can readily be traced with the 21cm hyperfine spin-flip transition and the Gunn-Peterson effect. At high redshifts, $z>1.6$, neutral hydrogen is currently traced by ground-based observations of Gunn-Peterson absorption lines against bright background QSOs. Such observations demonstrate that the bulk of the neutral hydrogen is in the Damped Ly-$\\alpha$ systems with column densities exceeding $2\\times10^{20}$ cm$^{-2}$. However, interpretation of these observations tends hampered by insufficient lines of sight and serious uncertainties associated with dust obscuration, gravitational lensing and intrinsic source size. These uncertainties lead to contradictory results from measurements associated with bright QSOs, faint QSOs and Gamma Ray Bursts \\cite{Jas05} \\cite{Pro06} \\cite{Por07}. However, with the SKA, galaxies will be detected at similar redshifts in 21cm line emission, which will lead to a clearer understanding of the distribution of gas in the Universe, and the manner in which the gas content of galaxies evolves with time. In order to measure the gas content of galaxies at the highest redshift, a deep pencil beam survey with the SKA is proposed. Currently proposed SKA specifications \\cite{ska07} are adopted, and used to generate artificial galaxy catalogues which are used to simulate the accuracy with which one simple parameter can be recovered, namely the cosmic HI density -- the comoving volume density of neutral hydrogen. ", "conclusions": "This simulation has demonstrated the large numbers of high-redshift galaxies that can be detected by the SKA in a significant, but feasible, HI survey of galaxies, and has demonstrated the high accuracy with which the cosmic HI density can be measured. Whilst galaxy numbers are low at redshifts below 1.8, this is largely due to the small field of view available to single pixel technology. Adoption of widefield detection technology has the potential to greatly increase numbers. Although the subsequent reduction of errors associated with shot noise and cosmic variance is important for many science goals, greater gains in the field of galaxy evolution study are likely to be made with better instantaneous sensitivity at lower frequencies. This will increase the ability to detect changes in the HI mass function and reduce the density extrapolation required to account for low-mass galaxies at redshifts approaching 3. An appropriate goal of $A/T>2\\times 10^4$ m$^2$ K$^{-1}$ is therefore suggested for frequencies below 500 MHz." }, "0801/0801.4517_arXiv.txt": { "abstract": "Models with dark energy decaying into dark matter have been proposed in Cosmology to solve the coincidence problem. We study the effect of such coupling on the cosmic microwave background temperature anisotropies. The interaction changes the rate of evolution of the metric potentials, the growth rate of matter density perturbations and modifies the Integrated Sachs-Wolfe component of cosmic microwave background temperature anisotropies, enhancing the effect. Cross-correlation of galaxy catalogs with CMB maps provides a model independent test to constrain the interaction. We particularize our analysis for a specific interacting model and show that galaxy catalogs with median redshifts $z_m=0.1-0.9$ can rule out models with an interaction parameter strength of $c^2\\simeq 0.1$ better than 99.95\\% confidence level. Values of $c^2\\le 0.01$ are compatible with the data and may account for the possible discrepancy between the fraction of dark energy derived from WMAP 3yr data and the fraction obtained from the ISW effect. Measuring the fraction of dark energy by these two methods could provide evidence of an interaction. ", "introduction": "Measurements of luminosity distances using supernovae type Ia (SNIa) \\cite{riess}, of the cosmic microwave background (CMB) temperature anisotropies with the WMAP satellite \\cite{wmap3}, large scale structure \\cite{lss,cole}, the integrated Sachs--Wolfe effect \\cite{fosalba-gazta2003,boughn-crittenden}, and weak lensing \\cite{weakl}, indicate that the Universe has entered a phase of accelerated expansion -see \\cite{reviews} for recent reviews. This acceleration is explained in terms of an unknown and nearly unclustered matter component of negative pressure, dubbed ``dark energy\" (energy density, $\\rho_{x}$), that currently contributes with about $75\\%$ of the total density. The remaining $25\\%$ is shared between cold dark matter ($\\rho_{c} \\sim 21\\%$), baryons ($\\rho_{b} \\sim 4\\%$), and negligible amounts of relativistic particles (photons and neutrinos). The pressureless non-relativistic matter component redshifts faster with expansion than the dark energy, giving rise to the so-called `coincidence problem' that seriously affects many models of late acceleration, particularly the $\\Lambda$CDM model: ``Why are matter and dark energy densities of the same order precisely today?\" \\cite{steinhardt}. To address this problem within general relativity one must either accept an evolving dark energy field or adopt an incredibly tiny cosmological constant and admit that the ``coincidence\" is just that, a coincidence that might be alleviated by turning to anthropic ideas \\cite{anthropic}. (It should be mentioned, however, the existence of proposals in which a vacuum energy density of about the right order of magnitude arises from the Casimir effect at cosmic scales -see \\cite{emili} and references therein). One way to address the coincidence problem, within general relativity, is to assume an interaction (coupling) between the dark energy component and cold dark matter such that the ratio $r \\equiv \\rho_{c}/\\rho_{x}$ evolves from a constant but unstable value at early times (in the radiation and matter dominated epochs) to a lower, constant and stable attractor at late times well in the accelerated expansion era \\cite{iqm0,iqm}. Before the late accelerated expansion was observed, Wetterich introduced interacting quintessence models to reduce the theoretical high value of the cosmological constant \\cite{wetterich}. Later, these kind of models were rediscovered, sometimes (but not always) in connection with the coincidence problem -see, e.g. \\cite{in-connection}. Other solutions (known as $f(R)$ models) require the Einstein-Hilbert action to be modified \\cite{f(R)}. Interacting quintessence models are testable since they predict differences in the expansion rates of the Universe, the growth of matter density perturbations, the Cosmic Microwave Background (CMB) temperature anisotropies and in other observables. In this paper we shall demonstrate that, as the Integrated Sachs-Wolfe effect measures directly the growth rate of matter density perturbations, can be used to detect variations on the evolution of the large scale structure with respect to the prediction of the concordance $\\Lambda$CDM model. As a toy model, we shall consider a spatially flat Friedmann--Robertson--Walker universe filled with radiation, baryons, dark matter and dark energy such that the last two components interact with each other through a coupling term, $Q = 3 {\\cal H} c^{2} (\\rho_{c} +\\rho_{x})$. Thus, the energy balance equations for dark matter and dark energy take the form \\\\ \\begin{equation} \\dot{\\rho}_{c}+3{\\cal H} \\rho_{c} = 3 {\\cal H} c^{2} (\\rho_{c} +\\rho_{x})\\, ,\\qquad \\dot{\\rho}_{x}+3 {\\cal H} (1+w_{x})\\rho_x= -3{\\cal H} c^{2} (\\rho_{c} +\\rho_{x}) \\, , \\label{balance} \\end{equation} \\\\ where $w_{x} = p_{x}/\\rho_{x} < -1/3$ is the equation of state parameter of the dark energy component and $c^{2}$ is a constant parameter that measures the strength of the interaction (it should not be confused with the speed of light that we set equal to unity). Derivatives are taken with respect to the conformal time, and ${\\cal H}= \\dot a/a$, with $a$ the scale factor of the Friedmann-Robertson-Walker metric. To satisfy the severe constraints imposed by local gravity experiments \\cite{peebles_rmph, hagiwara}, baryons and radiation couple to the other two energy components only through gravity. Our ansatz for $Q$ guarantees that the ratio between energy densities, $r$, tends to a fixed, attractor, value at late times. It yields a constant but unstable ratio at early times. The details of the calculation can be found in \\cite{iqm,olivares1,olivares2,olivares3}. This result holds irrespective of whether the dark energy is a quintessence (i.e., $ -1 2/3$, the mass is dominated by particles within $r$ (see \\citep{fillmoregoldreich} for further details). } \\label{fig:trajectory} \\end{figure*} \\begin{figure*} \\centering{ \\includegraphics[width=\\columnwidth]{density-MW-pruned} {\\hspace{0.2cm}} \\includegraphics[width=\\columnwidth]{density-numeric-analytic-pruned}} \\caption{ The left panel features the total numerically-obtained density for $\\epsilon=0.2$ [solid black curve, using expression~(\\ref{eq:total density})]. This solution is compared in the right panel to our analytic fit [dashed red curve, using expression~(\\ref{eq:rho_total})]. } \\label{fig:total density} \\end{figure*} ", "conclusions": "\\label{sec:conclusions} We have studied whether DM caustics in the halo of the Milky Way can amplify the flux of cosmic ray antiprotons and positrons which is received at the Earth. We have used the secondary infall model for our halo which naturally includes the caustics and assumed that the galactic DM is made of weakly interacting massive species. We have then taken into account the smearing of the caustics due to a present-day velocity dispersion of these particles of $0.03$~cm~s$^{-1}$. The cosmic ray antiprotons and positrons that are detected at the Earth originate from a region whose typical size is much larger than the shell thickness or even the shell separation. The associated horizon probes a large portion of the Milky Way and the coarse-grained average density~(\\ref{DM_shell_QD}) provides an adequate description. In the solar neighbourhood, the coarse-grained caustic density is the same as the smooth NFW or isothermal cored distributions usually assumed in the literature. The difference lies at the centre of the Milky Way. The lower-envelope density diverges there with an index of $\\sim 2.11$, hence a DM profile steeper than even in the NFW case. The coarse-grained shell density is assumed to be constant inside a sphere of radius $r_{\\rm cut-off}$ whose value is unknown. The smaller this cut-off radius, the more abundant DM at the galactic centre and consequently the stronger its signal should it reach the Earth. We have then computed the antiproton flux which the coarse-grained shell density yields at the Earth. The reach of the antiproton sphere depends on the cosmic ray propagation model but is always of order a few kiloparsecs. The MIN configuration is associated to a rather small antiproton range. The associated signal is not different from what is derived assuming NFW or isothermal cored distributions. The MAX propagation model is characterized on the contrary by efficient diffusion taking over a moderate galactic convection. The antiproton sphere probes the inner and denser regions of the Milky Way. We find that the antiproton signal is enhanced by a factor of $\\sim 30$ should the conventional smooth NFW profile be replaced by the coarse-grained shell density~(\\ref{DM_shell_QD}) for which a cut-off radius of 300 pc has been assumed. The MED set of propagation parameters corresponds to the best fit to the B/C data and features the intermediate situation. It leads to the exciting possibility that the antiproton signal is only boosted above 10 GeV in the presence of caustics. Depending on the WIMP annihilation cross section, the antiproton flux could be severely distorted at high energy as shown in Fig.~\\ref{fig:flux_WINO} where no artificial boost factor is required. Thus a promising window opens up around a few hundreds of Gev where future antiproton measurements are eagerly awaited. We are less optimistic for the positron signal. Depending on the cosmic ray propagation model, the positron flux at the Earth may be enhanced in the presence of shells with respect to a smooth NFW or isothermal cored DM profile. However, this situation arises only at low energy where the observations are already well explained by the sole secondary background component. Thus DM caustics cannot provide an explanation for the HEAT excess reported above $\\sim$ 10 GeV and consequently produced in the solar neighbourhood. The solution so far invoked is based on a WIMP with a hard positron annihilation spectrum like a Kaluza-Klein particle \\citep{Cheng:2002ej} or a neutralino with a dominant $W^{+}W^{-}$ channel \\citep{Hooper:2004bq,Delahaye:2007fr}. Boost factors of $\\sim$ 10 with respect to a smooth NFW DM halo are nevertheless necessary. Such a value is marginally possible \\citep{Lavalle_Maurin} in scenarios inspired by the $\\Lambda$-CDM N-body numerical simulations which point towards the existence of numerous and dense DM clumps inside which WIMP annihilation can be enhanced. We showed that the coarse-grained shell density~(\\ref{DM_shell_QD}) does not provide an alternative to explaining the HEAT excess. However, further data with higher precision and also a better understanding of the secondary positron background are needed to firmly exclude caustics as a possible explanation of the HEAT measurements. It remains very unlikely nevertheless that the present spectral form could be reproduced by caustics, at least by its coarse-grained distribution~(\\ref{DM_shell_QD}). We finally explored the possibility that the fine-grained distribution~(\\ref{fine_grained}) could yield a strong positron flux should the Earth be embedded inside the densest part of a caustic, a region where the DM density reaches its peak value $\\rho_{k,{\\sf max}}$. At high energy, the positron sphere shrinks. Averaging the fine-grained shell density~(\\ref{eq:density-caustic}) by its coarse-grained approximation~(\\ref{DM_shell_QD}) is no longer possible. Positrons that are received at the Earth have short diffusion lengths $\\lD$ and hence can be excellent tracers of nearby caustics. We investigated here the case of a positron line. Positrons that are detected exactly at the line energy $E = E_{S} \\equiv m_{\\chi}$ have vanishing diffusion length and if the Earth sits exactly inside the shell, an enhancement of the positron signal by a factor of $\\sim 13,000$ is naively expected. However, energy is measured with a limited accuracy and the energy bin of the line has a non-vanishing width. Averaging $\\lD$ over the line bin leads to a diffusion length which is still far larger than the typical caustic thickness or even separation. Our study showed no difference between the results of Section~\\ref{sec:cr_signal} and those derived with the fine-grained distribution~(\\ref{fine_grained}) once energy is properly averaged. We hypothesize that an extremely high energy resolution, presently unavailable, would allow to detect the nearby caustics. A word of caution is necessary at this stage though. The analysis of Section~\\ref{sec:draperies} is based on the assumption that cosmic rays diffuse on the inhomogeneities of the galactic magnetic field. The mean free path $\\lambda_{\\rm free}$ of their random walk may be derived from the space diffusion coefficient $K$ through the canonical relation \\be K(E) \\equiv {\\displaystyle \\frac{1}{3}} \\, \\lambda_{\\rm free} \\, \\beta \\;\\; . \\ee In this diffusion scheme and with the MED set of parameters, positrons with energy $\\epsilon$ cover on average a distance \\be \\lambda_{\\rm free} = 1.1 \\times 10^{-4} \\; {\\rm kpc} \\; \\epsilon^{0.7} \\label{lfree_close_line} \\ee before their next scattering on Alfv\\'en waves. Our treatment of cosmic ray propagation is definitely supported by the fact that $\\lambda_{\\rm free}$ is much smaller than the horizon size set by $\\lD$ and equation~~(\\ref{lD_close_line}). However, if the nearest DM caustic is very close to the Earth and lies at a distance which does not exceed $\\lambda_{\\rm free}$ or if the Earth is embedded inside the densest part of a shell, the diffusion hypothesis breaks down. Positrons will essentially drift along the lines of the magnetic field without encountering on their way any obstacle. Depending on the relative orientation of the local magnetic field with respect to the Earth and its nearest caustic, we could be possibly exposed to an intense high-energy positron flux whose evaluation is clearly beyond the scope of this article. {\\small Acknowledgment~: We thank Pierre Brun and Julien Lavalle for having provided us with a few typical positron and antiproton spectra arising from WIMP annihilation. RM thanks Sergei Shandarin for contributions, Niayesh Afshordi, Ed Bertschinger and Mike Kuhlen for discussions and LAPTH Annecy for hospitality and French programms PNC \\& ANR (OTARIE) for travel grants. Special thanks are due to Mark Vogelsberger for a careful reading of the manuscript and many useful comments and corrections. }" }, "0801/0801.1184.txt": { "abstract": "% context heading (optional) {} % aims heading (mandatory) {The aim of this work is to investigate the physical, structural and evolutionary properties of old, passive galaxies at $z>1.4$ and to place new constraints on massive galaxy formation and evolution.} % methods heading (mandatory) {We combine ultradeep optical spectroscopy from the GMASS project (\\emph{Galaxy Mass Assembly ultradeep Spectroscopic Survey}) with GOODS multi-band (optical to mid--infrared) photometry and HST imaging to study a sample of spectroscopically identified passive galaxies at $1.392$. No X-ray emission was found neither from individual galaxies nor from a stacking analysis of the sample. Only one galaxy shows a marginal detection at 24$\\mu$m. These galaxies have morphologies that are predominantly compact and spheroidal. However, their sizes ($R_e \\lesssim$ 1 kpc) are much smaller than those of spheroids in the present--day Universe. Their stellar mass surface densities are consequently higher by $\\approx$1 dex if compared to spheroids at $z\\approx0$ with the same mass. Their rest-frame $B$-band surface brightness scales with the effective radius, but the offset with respect to the surface brightness of the local Kormendy relation is too large to be explained by simple passive evolution. At $z\\approx1$, a larger fraction of passive galaxies follows the $z\\approx0$ size -- mass relation. Superdense relics with $R_e \\approx$ 1 kpc are extremely rare at $z\\approx0$ with respect to $z>1$, and absent if $R_e<1$ kpc. Because of the similar sizes and mass densities, we suggest that the superdense passive galaxies at $12$. The results are compared with theoretical models and the main implications discussed in the framework of massive galaxy formation and evolution. } % conclusions heading (optional), leave it empty if necessary {} ", "introduction": "Deep surveys provide the observational constraints needed to understand galaxy formation and evolution. In particular, many studies have been focused on massive galaxies (i.e. stellar mass ${\\cal M} > 10^{11}$ M$_{\\odot}$) as cosmological probes of the history of galaxy mass assembly. However, despite the remarkable success in finding and studying massive galaxies over a wide range of cosmic time, the global picture is far from being clear. Thanks to their simple and homogeneous properties (morphology, colors, passively evolving stellar populations, scaling relations) and being the most massive galaxies in the present-day Universe, early-type galaxies (ETGs) are crucial to investigate the cosmic history of massive galaxies (e.g. \\cite{alvio} and references therein). At $z<1$, the most recent results seem now to agree in indicating that the majority of massive ETGs (${\\cal M} > 10^{11}$ M$_{\\odot}$) were already in place at z$\\approx$ 0.7-0.8, with a number density consistent with the one at z=0, whereas the evolution is more pronounced for the lower mass ETGs (e.g. \\cite{fon04,yamada,bundy06,cdr06,borch,scarlata,brown,bundy07}). This mass--dependent evolutionary scenario, known as \"downsizing\" (\\cite{cowie,gavazzi}), was proposed to explain the galaxy star formation histories, i.e. with massive galaxies forming their stars earlier and faster than the low mass ones. Recent results suggest that the downsizing concept should extend also to the stellar mass assembly evolution itself, i.e. with massive galaxies assemblying their mass earlier (e.g. \\cite{cdr06,bundy06,franceschini, bundy07,perez}), thus providing new and stringent constraints for the current models of galaxy formation (e.g. \\cite{delucia}). The above results trace the evolution of the number density, luminosity and mass, but do not explain what is the mechanism with which ETGs assemble their mass and shape their morphology. Dissipationless ETG--ETG major merging (also called \"dry\" merging) has been advocated as an important mechanism to build up the masses of ETGs at $01$. The few ETGs identified spectroscopically so far up to $z\\approx 2.5$ are very red ($R-K>5-6$), dominated by passively evolving old stars with ages of 1-4 Gyr, have stellar masses typically ${\\cal M} \\gtrsim 10^{11} M_{\\odot}$, and are strongly clustered with $r_0 \\approx$ 8--10 Mpc (\\cite{cim04,glaze,mcc04,daddi05a,sar05,kong,kriek}). Daddi et al. (2005) were the first to realize that a large fraction of these ETGs have smaller sizes ($R_e \\lesssim 1$ kpc) (see also \\cite{cassata05}) and higher mass internal densities than present--day ETGs. This result was soon confirmed by other observations (\\cite{trujillo06,zirm07,longhetti07,toft07, trujillo07}). However, it is still unclear how to explain such size -- mass -- density properties in the context of ETG evolution. The existence of a substantial population of old, massive, passively evolving ETGs up to $z\\approx 2$ was not predicted in galaxy formation models available in 2004-2005, and opened the question on how it was possible to assemble such systems when the Universe was so young. A promising mechanism which can provide a better agreement with the observations seems to be the \"quenching\" of the star formation at high redshifts with AGN \"feedback\" (e.g. \\cite{gra,menci06}). The stellar ages and masses of the passive ETGs at $z\\approx1-2$ require precursors characterized by strong ($>100$ M$_{\\odot}$/yr) and short-lived (0.1-0.3 Gyr) starbursts occurring at $z > 2-3$. In addition, such precursors should also have a large clustering correlation length $r_0$ comparable to that of passive galaxies at lower redshifts ($z<2$) (e.g. \\cite{daddi00,firth02,kong,farrah06}), and compatible with that expected in the $\\Lambda$CDM models for galaxies located in massive dark matter halos and strongly biased environments. Examples of precursor candidates have been found amongst starburst galaxies selected at $z>2$ with a variety of techniques (e.g. $BzK$, \\cite{daddi04}; submm/mm, \\cite{chapman}; ``Distant Red Galaxies'', \\cite{franx}; optically-selected systems with high luminosity, \\cite{sha}; IRAC Extremely Red Objects, IEROs, \\cite{yan}; HyperEROs, \\cite{totani}, and ULIRGs at $z\\sim 1-3$ selected with {\\it Spitzer Space Telescope} \\cite{berta07}). Deep integral-field near-IR spectroscopy is being used to perform detailed studies of these precursor candidates in order to understand what are the main mechanisms capable to assemble massive galaxies with short timescales (e.g. \\cite{swin,nfs06,wright,law}). To date, the most detailed case is represented by a $BzK$--selected star-forming galaxy at $z=2.38$ which shows a massive rotating disk with high velocity dispersion which may become unstable and lead to the rapid formation of a massive spheroid (\\cite{genzel}). In this paper, we exploit the combination of GMASS ultradeep spectroscopy, HST imaging and optical -- to -- mid-infrared photometry to study the global properties of a new sample of passive galaxies at $1.310^{42}$ erg/s) is absent or is very heavily obscured. However, the possibility of dust obscuration seems unlikely because none of the galaxies has been detected at 24$\\mu$m, with only one marginal exception of one galaxy at $z=1.61$. $\\bullet$ The stellar masses, estimated through the photometric SED fitting, are in the range of $10^{10-11}$ M$_{\\odot}$ and the specific star formation rates are very low ($\\lesssim 3 \\times 10^{-2}$ Gyr$^{-1}$). The stellar masses estimated with model spectra including TP-AGB stars are systematically lower by 0.1--0.2 dex than those estimated with models which do not include this phase of stellar evolution. $\\bullet$ The HST+ACS morphological and surface brightness profile analysis indicate that the majority of the spectroscopically--selected passive galaxies have spheroidal morphologies consistent with being analogous to present-day ETGs. However, their sizes are smaller by a factor of $\\approx$2-3 than at $z\\approx0$, and imply that the stellar mass surface and volume internal densities are up to $\\approx$10 and $\\approx$ 30 times larger respectively. If literature data are added to the GMASS sample, we find that only a few passive systems at $1.21.3$. The ETGs at $z\\approx1$ which have the largest offsets with respect to the $z\\sim0$ size -- mass relation are the ones having the highest internal velocity dispersion, as expected from the ETG scaling relations. We find a hint that, for a fixed redshift $z\\approx$1, ETGs located within massive clusters are more preferentially located within the $z\\sim0$ size -- mass relation than ETGs located in lower density environments. $\\bullet$ Superdense massive ETGs with $R_e \\approx$ 1 kpc are extremely rare at $z\\approx0$ with respect to $z>1$, and absent if $R_e<1$ kpc. However, it might be possible that compact and dense remnants are \"hidden\" inside present-day ETGs if the size of ETGs grew through mechanisms such as the \"smooth envelope accretion\". $\\bullet$ Submillimeter--selected galaxies are the only systems at $z\\gtrsim$2 with sizes and mass surface densities (in gas) similar to those of the passive galaxies at $z\\approx$1--2. This suggests that a strong evolutionary link is present between these two galaxy populations. $\\bullet$ It is currently unclear how the possible link between SMGs and compact passive galaxies fits within a more general framework which takes into account also the other galaxy populations so far identified at $1$ 10$^{24}$ cm$^{-2}$. The amplitude of the reflected component may exceed 10\\% of the central unobscured luminosity. This is higher than the reflected fraction, of a few percent, observed in other Seyfert 2 sources like NGC 4945. We observe an emission line at 6.7 keV, possibly due to FeXXV, undetected in previous Chandra observations. The absorption column density associated with this line is less than 10$^{23}$ cm$^{-2}$, lower than the obscuration of the central source. We hypothesize that this highly ionized Fe line emission originates in warm gas, also responsible for a scattered component of continuum emission that may dominate the spectrum between 1 and 3 keV. We compare X-ray and maser emission characteristics of IC 2560 and other AGN that exhibit water maser emission originating in disk structures around central engines. The temperature for the region of the disk associated with maser action is consistent with the expected 400-1000K range. The clumpiness of disk structures (inferred from the maser distribution) may depend on the unobscured luminosities of the central engines. ", "introduction": "IC2560 is a nearby spiral ($cz\\sim 2925$\\,km\\,s$^{-1}$) that is classified as a Seyfert 2 \\citep{fai86} and known to host 22\\,GHz H$_2$O maser emission originating in an accretion disk around the central engine \\citep{ishi01}. Here we present analysis of a deep (archival) XMM-Newton EPIC observation of this active galactic nucleus (AGN). Previous observations of IC 2560 in the X-ray band using ASCA GIS \\citep{ishi01} and Chandra ACIS-S \\citep{iwa02,mad06} both pointed to a reflection dominated spectrum and high obscuring column density. IC 2560 is a relatively weak X-ray source, with an observed 2-10 keV flux of 3.3$\\times$10$^{-13}$ erg s$^{-1}$ cm$^{-2}$ that originates predominantly from a central source \\citep{mad06}. Interpretation of early GIS and ACIS-S observations was constrained by relatively limited sensitivity and energy resolution, in particular above 6 keV, where the bulk of the emission due to nuclear activity escapes into our line of sight. The EPIC instrument provides $\\sim$5$\\times$ larger collecting area at 6 keV and $\\sim$20\\% finer energy resolution (160 eV) than ACIS-S (chip 3). Although dominated by the central point source, flux detected with EPIC (1.1 kpc pixel$^{-1}$ at 1 keV) will include contribution from extended and off-nuclear emission components. As such, we note that the nuclear structure in IC 2560 appears to be relatively complex, as may be anticipated for a type-2 object. Hubble Space Telescope ACS/HRC observations at 3300 \\AA\\ detect patchy emission with multiple peaks within the central $\\sim$ 200 pc \\citep{mun07}, presumably due to dust. Relatedly, population synthesis models fit to optical spectra, obtained with the ESO 1.5m (3470-5450 \\AA) for the central $\\sim$ 300 pc, are suggestive of active star formation during the last 1.5-10 Myr \\citep{fer04}. Ground-based spectroscopic study has also detected asymmetry in the [OIII] line profile (a blue excess) that may be indicative of outflow at $\\sim$ 100 km s$^{-1}$ projected on the sky plane in a 300$\\times$400 pc central patch \\citep{schu03}. Because the inclination of the galaxy is 63$^\\circ$, and the inclination of the central engine (as inferred from maser emission) may be even closer to edge-on, the actual flow speed may be larger. The observed H$_2$O maser spectrum exhibits emission close to the systemic velocity as well as blue and red-shifted complexes, offset, approximately symmetrically, by $\\pm$ 200 - 420 km s$^{-1}$ \\citep{bra03}. This pattern is an indicator of emission from a rotating edge-on disk structure (e.g., NGC 4258; \\cite{nak93}). \\cite{ishi01} reported line-of-sight accelerations for the ``systemic emission'' and mapped its emission distribution using Very Long Baseline Interferometry (VLBI). Based on the assumption that the emission traced a thin, rotating, circular disk that is observed edge-on, they used the velocities of high-velocity emission to estimate disk radius and dynamical mass. Follow-up VLBI observation by Greenhill \\etal (2008, in preparation) enabled mapping of both systemic and high-velocity emission, thus tracing the disk structure, demonstrating Keplerian rotation of material around the central engine in a relatively thin distribution, and enabling more certain modeling of the accretion disk geometry and estimation of the central mass. The observed molecular gas lies at radii 0.08 - 0.27 pc and the inferred enclosed mass is 2.9$\\times$10$^6$M$_\\odot$ with $\\sim$ 20\\% accuracy. The thinness of disk (h/r $\\ll$1) as well as presence of H$_2$O emission suggest disk material that is relatively cool ($\\la$10$^3$ K). Due to generally high extinction toward Seyfert-2 central engines, X-ray and radio observations together are especially useful in the study of the physical processes that occur at radii $\\la$1 pc. Measurement of X-ray spectra enable estimation of intrinsic luminosity, obscuring column density, temperature and abundance. Time monitoring facilitates estimation of size scales for high-energy phenomenon. VLBI observations of H$_2$O maser emission enable mapping of accretion disk geometry and orientation to the line of sight, and estimation of orbital radii for molecular material, central engine mass and Eddington luminosity (\\cite{gre07} and references therein). The presence of maser emission implies the existence of a reservoir of relatively cool material and at the same time, a heating mechanism to maintain population inversion. The necessary energy may be imparted by viscous heating within the disk (e.g., \\citep{des98}) or external irradiation \\citep{neu94,wat94}. However, in either case, survival of molecular gas demands a shielding column from UV and X-ray emission generated by, e.g., the central engine and coronae. This column may arise in outflowing material, disk surface layers, or disk material at small radii, thus providing additional quantitative constraints on overall structure around the central engine and energetics. In section 2, we discuss the data reduction of the EPIC data, while section 3 focuses on analysis of the source spectrum. Estimation of the intrinsic emission of the central engine (i.e., corrected for absorption along line of sight) is discussed in section 4. X-ray properties of IC 2560 and other extragalactic H$_2$O maser sources believed to originate in accretion disks are considered in section 5; there we put IC 2560 into context and we discuss possible trends that may indicate connection between characteristics estimated from VLBI observations of masers and X-ray observations of their host AGN. Throughout the paper we adopt H$_{0}$=70 km s$^{-1}$ Mpc$^{-1}$. ", "conclusions": "A deep 70 ks XMM-Newton spectrum for IC 2560 shows evidence for a warm scattered continuum component as well as number of emission lines not detected in earlier Chandra observations. These include K$\\alpha$ emission from Mg (1.25 keV), Ca (3.6 keV) and Ni (7.47 keV) as well as Fe K$\\beta$ (7.05 keV) and K$\\alpha$ line emission from Fe XXV (6.7 keV). We do not detect line emission from species at intermediate ionization states, between the nearly neutral material that gives rise to the K$\\alpha$ emission, and the warm scattering medium with FeXXV. The direct X-ray emission from the central source appears to be completely obscured up to 10 keV; the inferred column density is in excess of a few times 10$^{24}$ cm$^{-2}$. This is in agreement with a 25\\% upper limit on the magnitude of the Compton shoulder. The reflecting medium is mostly neutral, optically thick and could also provide the obscuring column toward the central engine. The 6.7 keV line is associated with the optically thin, mostly ionized, scattering medium, with significantly lower density. The dominance of the resonance line at 6.7 keV indicates the density has to be lower than 10$^{23}$ cm$^{-2}$, while the estimated fraction of intrinsic emission scattered along line-of-sight indicates a lower limit of 10$^{22}$ cm$^{-2}$. The scattering medium can plausibly exist on the scale of a few parsec, but the exact geometry is unknown. Limits on density of the scatterer indicate a covering fraction between 0.1 and 1. Using a small sample of H$_{2}$O disk maser hosts, selected using physically motivated criteria, we find that the unobscured luminosity of the central source may shape molecular disk structure in this sample. The inner radius as well as degree of substructure in the disk may be related to the unobscured X-ray emission, though a larger sample is required to confirm this trend." }, "0801/0801.1662_arXiv.txt": { "abstract": "We construct a little Higgs model with the most minimal extension of the standard model gauge group by an extra $U(1)$ gauge symmetry. For specific charge assignments of scalars, an approximate $U(3)$ global symmetry appears in the cutoff-squared scalar mass terms generated from gauge bosons at one-loop level. Hence, the Higgs boson, identified as a pseudo-Goldstone boson of the broken global symmetry, has its mass radiatively protected up to scales of 5-10~TeV. In this model, a $Z_2$ symmetry, ensuring the two $U(1)$ gauge groups to be identical, also makes the extra massive neutral gauge boson stable and a viable dark matter candidate with a promising prospect of direct detection. ", "introduction": " ", "conclusions": "" }, "0801/0801.3667_arXiv.txt": { "abstract": "Through the modelling of the Spectral Energy Distribution of blazars we can infer the physical parameters required to originate the flux we see. Then we can estimate the power of blazar jets in the form of matter and fields. These estimate are rather robust for all classes of blazars, although they are in part dependent of the chosen model (i.e. leptonic rather than adronic). The indication is that, in almost all cases, the carried Poynting flux is not dominant, while protons should carry most of the power. In emission line blazars the jet has a comparable, and often larger, power that the luminosity of the accretion disk. This is even more true for line--less BL Lacs. If the jet is structured at the sub--pc scale, with a fast spine surrounded by a slower layer, then one component sees the radiation of the other boosted, and this interplay enhances the Inverse Compton flux of both. Since the layer emission is less beamed, it can be seen also at large viewing angles, making radio--galaxies very interesting GLAST candidates. Such structures need not be stable components, and can form and disappear rapidly. Ultrafast TeV variability is challenging all existing models, suggesting that at least parts of the jets are moving with large bulk Lorentz factors and at extremely small viewing angles. However, these fast ``bullets\" are not necessarily challenging our main ideas about the energetics and the composition of the bulk of the jet. ", "introduction": "EGRET, onboard {\\it CGRO}, and the ground based Cherenkov telescopes showed that blazars are the most powerful high energy extragalactic emitters, and allowed to know the bolometric luminosity of these objects. Now, just after the launch of {\\it AGILE} and just before {\\it GLAST}, we are preparing for (and already tasting) the possibility to have simultaneous data in the optical, X--rays and the GeV bands (and possibly the TeV one). This will be possible mainly by {\\it Swift}: its rapid slew and flexible scheduling will ensure good quality optical and X--ray data while the $\\gamma$--ray observations are still ongoing. What was an exception in the EGRET era, will be routine. This is therefore a risky and at the same time healthy time to put forward new ideas concerning blazars, that {\\it GLAST} can falsify. In this contribution, I will try to summarize how the power of jets can be derived, and what inferences can we draw from that. I will discuss the ultra--fast TeV variability recently observed in PKS 2155--304, arguing that, contrary to previous claims, it is unlikely that the jet of this source is magnetically dominated. Furthermore, I will point out that the fact that the TeV emitter BL Lac objects are also the least powerful blazars opens up the possibility to slow down their jets by the Compton rocket effect. \\begin{figure} \\vskip -0.2 true cm \\hskip -0.5 true cm \\centerline{\\psfig{figure=ghisellini_f1.ps,width=11truecm}} \\vskip -0.5 true cm \\caption{SEDs of 3C 454.3 at different epochs. Top panel: the SED in 2000, corresponding to the {\\it Beppo}SAX observations discussed in \\cite{tav02}. % The other points are not simultaneous (see \\cite{pian06} % and references therein). Mid panel: the SED during the huge optical flare in 2005, as described in \\cite{pian06, giommi06}. % Bottom panel: the SED on July 26, 2007, as observed by {\\it AGILE}\\cite{agile07}, and {\\it Swift}. The optical flux in the $R$ band comes from the Tuorla observatory. The optical and X--ray fluxes are corrected for Galactic extinction ($A_V=0.355$). The solid and dashed lines correspond to our modelling. The dotted line is the contribution from the accretion disk (assumed to be a simple black--body). From \\cite{gg07}. % } \\label{454} \\end{figure} ", "conclusions": "" }, "0801/0801.4876_arXiv.txt": { "abstract": "Multi-frequency (4.6, 5, 5.5, 8, 8.8, 13, 15, 22 \\& 43 GHz) polarization observations of 6 ``blazars'' were obtained on the American Very Long Baseline Array (VLBA) over a 24-hr period on 2 July 2006. Observing at several frequencies, separated by short and long intervals, enabled reliable determination of the distribution of Faraday Rotation on a range of scales. In all cases the magnitude of the RM increases in the higher frequency observations, implying that the electron density and/or magnetic field strength is increasing as we get closer to the central engine. After correcting for Faraday rotation, the polarization orientation in the jet is either parallel or perpendicular to the jet direction. A transverse Rotation Measure (RM) gradient was detected in the jet of 0954+658, providing evidence for the presence of a helical magnetic field surrounding the jet. For three of the sources (0954+658, 1418+546, 2200+420), the sign of the RM in the core region changes in different frequency-intervals, indicating that the line-of-sight component of the magnetic field is changing with distance from the base of the jet. We suggest an explanation for this in terms of bends in a relativistic jet surrounded by a helical magnetic field; where there is no clear evidence for pc-scale bends, the same effect can be explained by an accelerating/decelerating jet. ", "introduction": "AGN jets emit synchrotron radiation that can be detected at all radio frequencies. This radiation is often highly linearly polarized, which provides important information on the degree of order and orientation of the magnetic field in these jets. The type of AGN studied in this experiment are known as ``blazars'', which have jets pointed close to our line of sight (LoS) and often exhibit strong variability in total flux and linear polarization over a broad range of frequencies from $\\gamma$-ray to radio. We observed 6 sources with the American Very Long Baseline Array (VLBA) at 8 frequencies from 4.6 GHz to 43 GHz over a 24-hr period on 2 July 2006. This radio interferometer provides milliarcsecond resolution which corresponds to the parsec scale structure of these jets. Even though these AGN are intrinsically two-sided, pc-scale observations typically show a one sided ``core-jet'' structure due to Doppler boosting of the radiation from the relativistic jet pointed towards us. The jet moving away from us is usually too faint to be detected. The radiation detected from the jet is generally optically thin while the core region usually displays a flat spectrum attributed to synchrotron self-absorption. In the optically thin regime the polarization orientation is perpendicular to the magnetic field, while in the optically thick/flat spectrum region the polarization is parallel to the magnetic field. Recent MHD simulations have provided an almost complete explanation of how these jets are launched, accelerated and collimated close to the black hole, see \\cite{MeierJapan} and references therein. The global magnetic field structure expected on these scales is helical. It is possible that this structure changes when the flow gets disrupted by shocks after a few hundred Schwarzschild radii, but observational evidence of helical fields on scales larger than this \\cite{Asada2002, GabuzdaMurray2004, Mahmud2008} suggests that remnants of the earlier magnetic field structure remain or that a current driven helical kink instability is generated \\cite{NakamuraJapan, Carey2008}. ", "conclusions": "Several phenomenon, for example, the polarization structure \\cite{Lyutikov2005}, RM gradients \\cite{Asada2002} and circular polarization generation \\cite{GabCP2007}, can be understood by considering a helical magnetic field geometry for pc-scale AGN jets. In this paper, we consider how regions with different RM signs can also be explained within a helical magnetic field model, as places where the jet is observed at angles greater than or less than $1/\\Gamma$ due to bends in the jet or due to an accelerating/decelerating straight jet. (A longitudinal jet magnetic field with a change in the angle to the LoS could also cause a RM sign change, but this does not correspond to the observed magnetic field in 0954+658 or 2200+420.) It's important to note that VLBI resolution is usually not sufficient to completely resolve the true optically thick core, therefore, the VLBI ``core'' consists of emission from the true core and some of the optically thin inner jet. So if bends occur on scales smaller than the observed VLBI ``core'', core RMs with different signs can be derived from observations at different frequencies (ie. probing different scales of the inner jet). In our future work, we will attempt to reconstruct the 3-D path of the jet through space using the combined information from the observed distributions of the total intensity, linear polarization, spectral index and rotation measure." }, "0801/0801.1986_arXiv.txt": { "abstract": "The year 2007 has furnished us with outstanding results about the origin of the most energetic cosmic rays: a flux suppression as expected from the GZK-effect has been observed in the data of the HiRes and Auger experiments and correlations between the positions of nearby AGN and the arrival directions of trans-GZK events have been observed by the Pierre Auger Observatory. The latter finding marks the beginning of ultra high-energy cosmic ray astronomy and is considered a major breakthrough starting to shed first light onto the sources of the most extreme particles in nature. This report summarizes those observations and includes other major advances of the field, mostly presented at the 30$^{\\rm th}$ International Cosmic Ray Conference held in M\\'erida, Mexico, in July 2007. With increasing statistics becoming available from current and even terminated experiments, systematic differences amongst different experiments and techniques can be studied in detail which is hoped to improve our understanding of experimental techniques and their limitations. ", "introduction": "Understanding the origin of the highest energy cosmic rays is one of the most pressing questions of astroparticle physics. Cosmic rays with energies exceeding $10^{20}$ eV have been observed for more than 40 years (see e.g.\\ \\cite{Nagano-Watson}) but due to their low flux only some ten events of such high energies could be detected up to recently. There are no generally accepted source candidates known to be able to produce particles of such extreme energies. Moreover, there should be a steeping in the energy spectrum near $10^{20}$ eV due to the interaction of cosmic rays with the microwave background radiation (CMB). This Greisen-Zatsepin-Kuzmin (GZK) effect \\cite{GZK} severely limits the horizon from which particles in excess of $\\sim 6\\cdot10^{19}$ eV can be observed. For example, the sources of protons observed with $E\\ge 10^{20}$ eV need to be within a distance of less than 50 Mpc \\cite{Harari-06}. The non-observation of the GZK-effect in the data of the AGASA experiment \\cite{Takeda-03} has motivated an enormous number of theoretical and phenomenological models trying to explain the absence of the GZK-effect and has stimulated the field as a whole. Only this year, with the final analysis of the HiRes-data \\cite{Abbasi-07a} and the advent of high-statistics and high quality hybrid data from the Pierre Auger Observatory (PAO) \\cite{Yamamoto-07}, the situation has changed considerably: a suppression such as expected from the GZK-effect is now observed with high statistical significance. The very recent breaking news about the observation of directional correlations of the most energetic Pierre Auger events with the positions of nearby AGN \\cite{Abraham-07} complements the observation of the GZK effect very nicely and provides evidence for an astrophysical origin of the most energetic cosmic rays. Another key observable allowing to discriminate different models about the origin of high-energy cosmic rays is given by the mass composition of cosmic rays. Unfortunately, such measurements are much more difficult due to their strong dependence on hadronic interaction models. Only primary photons can be discriminated safely from protons and nuclei and recent upper limits to their flux largely rule out top-down models, originally invented to explain the apparent absence of the GZK-effect in AGASA data. In this article, prepared for the TAUP conference in Sendai (Japan), we describe the status of each of these topics, as reported during the recent International Cosmic Ray Conference held in M\\'erida, Mexico, in July 2007 (ICRC2007) and in publications becoming available since. ", "conclusions": "Remarkable progress has been made in cosmic ray physics at the highest energies, particularly by the start-up of the (still incomplete) Pierre Auger Observatory. The event statistics above $10^{19}$\\,eV available by now allows detailed comparisons between experiments and indicates relative shifts of their energy scales by $\\pm 25$\\,\\%. Given the experimental and theoretical difficulties in measuring and simulating extensive air showers at these extreme energies, this may be considered a great success. On the other hand, knowing about overall mismatches of the energy scales between experiments may tell us something. Clearly, in case of fluorescence detectors better measurements of the spectral and absolute fluorescence yields and their dependence on atmospheric parameters are needed and will hopefully become available in the very near future \\cite{fluorescence}. This should furnish all fluorescence experiments with a common set of data. Differences in the calibration between surface detectors and fluorescence telescopes, best probed by hybrid experiments like Auger and TA, may then be used to test the modelling of EAS. The muon component at ground, known to be very sensitive to hadronic interactions at high energies \\cite{Drescher-04}, could in this way serve to improve hadronic interaction models in an energy range not accessible at man-made accelerators. In fact, several studies (e.g.\\ \\cite{Engel-07}) indicate a deficit of muons by 30\\,\\% or more in interaction models like QGSJET. The energy scale is of great importance also for the AGN correlation discussed in the previous section. As shown in \\cite{AGN-long-07}, the correlation sets in abruptly at an (Auger) threshold energy of about 57 EeV. Already a downshift in energy by 17\\,\\% (the mismatch between Auger and HiRes) would weaken the signal by more than 3 orders of magnitude to make it basically disappear. Thus, verification of the correlation signal by HiRes or AGASA would need to be done for a threshold energy (on their scale) of 67 EeV and 85 EeV, respectively. In this energy range, HiRes observes a spectral slope of $\\gamma=5.1\\pm0.7$ \\cite{Bergman-07b}, i.e.\\ the number of events available for a correlation analysis would, according to Table \\ref{tab:expts}, drop to about 12 (HR-I) and 4 (HR-II) when taking the rise of the apertures into account. This would amount to about half the statistics of Auger, well in agreement with the quoted exposures. Unfortunately, the angular resolution of monocular reconstruction is by far too poor for such a test. Only stereo data could provide the required angular resolution. However, in this case the expected statistics of about 5 events above threshold (based upon the numbers and exposures given above) appears too small for any verification. In fact, the distance parameter of the correlation of 71 Mpc may indicate a mismatch of the energy scale: For protons above 57 EeV the GZK horizon would be 200 Mpc \\cite{Harari-06} but already for 20\\,\\% higher energy it would shrink by more than a factor of two to become consistent to the correlation parameter. Another puzzling feature is the observed small deflection of particles which suggests dominantly protons as primaries. Note that 90\\,\\% of the events (20/22) off the galactic plane are correlated to within $\\sim 3^\\circ$ which AGN positions which is very unlikely for heavy nuclei. On the other hand, the elongation curves in Fig.\\,\\ref{fig:Xmax} suggests an admixture of heavy nuclei by more than 10\\,\\%. This may be related again to imperfections of the hadronic interaction models used for comparison in Fig.\\,\\ref{fig:Xmax}. All of this tells us that the near future will be highly exciting: The question of the energy scales will soon be settled and more detailed comparisons between experiments will become possible. The shape of the energy spectrum in the GZK region will tell us about the source evolution, the composition in the ankle region will answer the question about the G-EG transition, observations of cosmogenic photons and neutrinos are in reach and in case of neutrinos will probe the GZK effect over larger volumes, the correlations will be done with better statistics, with improved search techniques and with more appropriate source catalogues and source selection parameters to tell us about source densities, and the true sources of EHECRs. Very important to note is that different pieces of information start to mesh and are being accessed from different observational techniques and can be cross-checked: {\\em The big picture is being painted!} Given the scientific importance of this, it would be a mistake to have only one observatory - even when operated as a hybrid detector - taking data. The TA project and its extensions will be very important particularly in the sub-GZK range but, unfortunately, will be too small to collect sufficient statistics at the highest energies. Auger-North will be imperative here and needs immediate vigorous support. The next generation experiment JEM EUSO to be mounted at the Exposed Facility of Japanese Experiment Module JEM EF will potentially reach much larger exposures but still faces many experimental challenges to be addressed. \\vspace*{-5mm} \\subsection*{Acknowledgement} I would like to thank the organizers of the excellent TAUP meeting in Sendai for the invitation to give a review talk at this exciting time. Also, its a pleasure to thank many of my colleagues for stimulating discussions. The German Ministry for Research and Education (BMBF) and the Deutsche Forschungsgemeinschaft (DFG) are gratefully acknowledged for financial support." }, "0801/0801.3818_arXiv.txt": { "abstract": "Simultaneous multiwavelength studies of X-ray binaries have been remarkably successful and resulted in improved physical constraints, a new understanding of the dependence of mass accretion rate on X-ray state, as well as insights on the time-dependent relationship between disk structure and mass-transfer rate. I will give some examples of the tremendous gains we have obtained in our understanding of XRBs by using multiwavelength observations. I will end with an appeal that while Spitzer cryogens are still available a special effort be put forth to obtaining coordinated observations with emphasis on the mid-infrared: Whereas the optical and near-IR originate as superpositions of the secondary star and of accretion processes, the mid-IR crucially detects jet synchrotron emission from NSs that is virtually immeasurable at other wavelengths. A further benefit of Spitzer observations is that mid-infrared wavelengths can easily penetrate regions that are heavily obscured. Many X-ray binaries lie in the Galactic plane and as such are often heavily obscured in the optical by interstellar extinction. The infrared component of the SED, vital to the study of jets and dust, can be provided {\\it only} by Spitzer; in the X-rays we currently have an unprecedented six satellites available and in the optical and radio dozens of ground-based facilities to complement the Spitzer observations. ", "introduction": "X-ray binaries (XRBs) owe their prominence to one of the most efficient energy release mechanisms known: accretion onto a compact object. The energy released by accretion can be spread over essentially the entire electromagnetic spectrum. Since each part of the spectrum provides distinct and often time-variable information (see, e.g., Fig. 1), attempts to understand these systems by concentrating on information from a limited wavelength range may lead to contradictory and misleading conclusions (see, e.g., Figs. 2-4). It is important to study them simultaneously over a broad range of the electromagnetic spectrum; spectral energy distributions (SEDS) are particularly useful because they clearly reveal multiple emission components and reflect the physics and geometry of the emitting regions. Multiwavelength studies have been remarkably successful and resulted in improved physical constraints on the systems, a new understanding of the dependence of mass accretion rate on X-ray state, as well as insights on the time-dependent relationship between disk structure and mass-transfer rate, (e.g., Vrtilek~\\etal~1990,1991; Hasinger~\\etal~1990; McClintock~\\etal~2001; Fuchs~\\etal~2003; Homan~\\etal~2005) One of the fundamental unsolved problems of accretion physics is how the outgoing power is distributed between electromagnetic radiation and mechanical power (relativistic jets), and what determines a switch between those two output channels. Relativistic jets are amongst the most energetic phenomena in the Universe. In our Galaxy, jets have been detected in young stellar objects, massive binaries, symbiotic stars, BH and NS XRBs, super-soft X-ray sources, and planetary nebulae. They are also believed to be responsible for $\\gamma$-ray bursts (Fender~\\etal~2004). Despite the tremendous range of scale, jets in stellar mass systems share physical properties with those in super-massive BHS. The most significant property shared is that {\\it all systems that produce jets also have accretion disks} (Livio 2002). This implies that jet energy is obtained through accretion power and that they may play a major role in the transport of angular momentum of the inflalling gas. As spatial separation is generally impossible owing to limitation in resolution, {\\it multiwavelength studies are essential to separate the stellar, wind, disk, and jet components. The broad band spectrum is also necessary to test theoretical models for jet formation and to constrain fundamental parameters of jet physics}. \\begin{figure} \\centering { \\scalebox{0.22}{\\includegraphics{vrtilek_s_fig1.ps}} } { \\scalebox{0.30}{\\includegraphics{vrtilek_s_fig2.ps}} } \\caption{ LEFT: Broadband energy spectra of the black-hole binary GRO~J1655-40 illustrating variability within one source. Each color represents simultaneous observations. (Migliari~\\etal~2007; see also Tomsick contribution to these proceedings). RIGHT: Broadband energy spectra of a selection of black hole and neutron star binaries showing variability between sources (ATCA and RXTE values are from a survey conducted by the Disk/Jet Consortium in June of 2006; values for Sco X-1 are from Wachter~\\etal~2006). } \\label{fig1sub} \\vspace{-3mm} \\end{figure} \\begin{figure} \\centering { \\scalebox{0.3}{\\includegraphics{vrtilek_s_fig3.ps}} } { \\scalebox{0.25}{\\includegraphics{vrtilek_s_fig4.ps}} } \\caption{ LEFT: RXTE ASM countrates (courtesy of the RXTE ASM team) of the black hole candidate Cygnus X-1 with typical countrates for ``Low\", ``Intermediate\", and ``High\" states indicated. RIGHT: Broadband energy spectrum of Cygnus X-1 (Gierlin'ski~\\etal~1993). It is clear that the designations ``High\" and ``Low\" states are based on a narrow energy range (2-10 keV) and are not valid when the full X-ray band is available. } \\label{fig2sub} \\vspace{-3mm} \\end{figure} \\begin{figure} \\centering { \\scalebox{0.23}{\\includegraphics{vrtilek_s_fig5.ps}} } { \\scalebox{0.32}{\\includegraphics{vrtilek_s_fig6.ps}} } \\caption{ LEFT: RXTE ASM countrates (courtesy of the RXTE ASM team) of the pulsing XRB Her X-1. Arrows mark the dates of the SEDs plotted on the right. RIGHT: Simultaneous SEDs of Her X-1 clearly show that the optical and ultraviolet flux (left side) remain largely unaffected at times when the extreme ultraviolet to hard X-ray flux (right side) takes a dramatic plunge. An indication that X-ray flux is not a dependable measure of mass accretion rate. Vrtilek~\\etal~2001). } \\label{fig2sub} \\vspace{-3mm} \\end{figure} \\begin{figure} \\centering { \\scalebox{0.18}{\\includegraphics{vrtilek_s_fig7.ps}} } \\hspace{1cm} { \\scalebox{0.35}{\\includegraphics{vrtilek_s_fig8.ps}} } \\caption{ LEFT: Hardness vs. intensity data from RXTE observations of the Galactic Z-source Sco X-1 (Barnard, Kolb, \\& Osborne 2003). RIGHT: UV continuum vs line flux in 1224-1986$\\AA$ band of Sco X-1 verifying that mass accretion rate increases from HB to NB to FB and does not correlate with X-ray flux. (Vrtilek~\\etal~1991).} \\label{cygscocyg} \\end{figure} ", "conclusions": "" }, "0801/0801.1506_arXiv.txt": { "abstract": "The scaling relations between rotation velocity, size and luminosity form a benchmark test for any theory of disk galaxy formation. We confront recent theoretical models of disk formation to a recent large compilation of such scaling relations. We stress the importance of achieving a fair comparison between models and observations. ", "introduction": "Understanding the origin and nature of galaxy scaling relations is a fundamental quest of any successful theory of galaxy formation. The success of a particular theory will be judged by its ability to predict the slope, scatter, and zero-point of any robust galaxy scaling relation at any particular wavelength. The scaling relations between rotation velocity, $V$, size, $R$, and luminosity, $L$, are of special interest as they are linked via the virial theorem. Courteau \\etal (2007; hereafter C07) compiled a sample of 1303 disk galaxies for which accurate rotational velocities and near-infrared sizes and luminosities are available. We refer the reader to C07 for details about this compilation. The $VLR$ scaling relations for these galaxies are shown in Fig.~1. The solid black lines in each panel show orthogonal linear fits to the combined data set, while the dashed lines show the $2\\sigma$ scatter in these relations. The points are color-coded according to central surface brightness. This reveals the fundamental independence of surface brightness of the $VL$ relation. As discussed in C07, the $VL$ relation is also independent of disk size and residuals from the size-luminosity relation. This is unexpected, as the simplest models of galaxies embedded in dark matter halos predict a strong surface brightness dependence, unless disk galaxies are dominated by dark matter within their optical regions (Courteau \\& Rix 1999; Dutton \\etal 2007). ", "conclusions": "" }, "0801/0801.4924_arXiv.txt": { "abstract": "We review recent 3D cosmological hydrodynamic simulations of primordial star formation from cosmological initial conditions (Pop III.1) and from initial conditions that have been altered by radiative feedback from stellar sources (Pop III.2). We concentrate on simulations that resolve the formation of the gravitationally unstable cloud cores in mini-halos over the mass range $10^5 < M/\\Msun < 10^7 $ and follow their evolution to densities of at least $10^{10} \\cmm3$ and length scales of $<10^{-2}$ pc such that accretion rates can be estimated. The advent of ensembles of such simulations exploring a variety of conditions permits us to assess the robustness of the standard model for Pop III.1 star formation and investigate scatter in their formation redshifts and accretion rates, thereby providing much needed information about the Pop III IMF. The simulations confirm the prediction that Pop III.1 stars were massive ($\\sim 100 \\Msun$), and form in isolation in primordial mini-halos. Simulations of Pop III.2 star forming in relic HII regions suggest somewhat lower masses ($\\sim 30 \\Msun$) which may help explain the chemical abundances of extremely metal poor stars. We note that no 3D simulation at present has achieved stellar density let alone followed the entire accretion history of the star in any scenario, and thus the IMF of Pop III stars remains poorly determined theoretically. ", "introduction": "Dark matter mini-halos in the mass range $10^5-10^6$ solar mass virializing at high redshift are believed to collect enough primordial gas within them to host the formation of the first generation of stars (Population III). Mediated by molecular hydrogen cooling, gas in the centers of these halos condenses and becomes unstable to gravitational fragmentation with a typical mass scale of a few hundred solar masses. Pioneering high resolution 3D hydrodynamic cosmological simulations by Abel et al. (2002) and Bromm et al. (2002), and more recently by Yoshida et al. (2006) and O'Shea \\& Norman (2007a) have shown that gravitational fragmentation produces only one massive fragment per halo. If this gas does not fragment further as it approaches stellar density, Pop III stars with this typical mass would be produced. The fates and feedback effects of such stars has many interesting consequences for the early structure formation as discussed elsewhere in this volume (see review by Ciardi). While this picture is now firmly established, based as it is on rather simple physics elucidated by the simulations, it is important to realize the remaining uncertainties. First, as of 2003 when the first reviews of primordial star formation were being written (Barkana \\& Loeb 2001, Bromm \\& Larson 2004, Glover 2005), a rather small number of such simulations had been done, raising the question of how ubiquitous this mode of star formation was in the early universe. Second, despite the large range of scales covered by the simulations, they were necessarily terminated well before stellar density was reached due to missing physics. The possibility of sub-fragmentation could not be ruled out, with consequent uncertainty on the mass scale of the first stars. Third, as emphasized by Barkana \\& Loeb (2004), for technical reasons the simulations were carried out in quite small cosmological volumes (< 1 Mpc comoving), the result being that the redshift of formation of the first stars found ($z\\sim 20$) was underestimated. Larger volumes would contain rarer peaks in the density field, and these would presumably form Pop III stars earlier (White \\& Springel 2000). Could the higher densities and temperatures of this earlier epoch alter the mass scale of the truly first stars? And finally, there was a host of complicated feedback effects--radiative, kinetic, and chemical--which were not included in these first simulations which could ``mess up\" the simple picture by altering the initial conditions and the cooling properties of the collapsing cloud cores. The most concerning of these feedback effects was the build-up of an FUV background by emissions from the first stars that could photo-dissociate the hydrogen molecules which mediate Pop III star formation in the first place (Haiman, Rees \\& Loeb 1997, Ciardi, Ferrara \\& Abel 2000, Haiman, Abel \\& Rees 2001). The first simulations examining this ``negative feedback effect\" (Machacek, Bryan \\& Abel 2001, hereafter MBA; and Yoshida et al 2003, hereafter Y03) concluded that Pop III star formation would not be suppressed, but merely delayed. The concept of a critical halo mass capable of forming a Pop III star was introduced by MBA. They found that the critical halo mass is an increasing function of the FUV background mean intensity in the Lyman-Werner bands (11.2-13.6 eV) suggesting that Pop III star formation would become self-limiting. Y03 found that gas cooling becomes inefficient above $J_{21}=0.01$ due to the photo-destruction of \\h2. Here $J_{21}$ is the mean intensity of the FUV background in units of $10^{-21}$ ergs/cm$^2$/sec/ster/Hz. Using this information Y03 constructed a semi-analytic model of the cosmic history of Pop III star formation. They found a rapidly rising population of Pop III stars over the redshift interval 35 > z > 20 ``the rise of Pop III\", which begins to become self-regulated at $z \\sim 25$ when the FUV background attains $J_{21}=0.01$. They did not consider the lower redshift evolution as this was out of the range of their simulations, but rather speculated that Pop II star formation dominated the cosmic star formation history below $z \\sim 20$. It is important to point out that neither the MBA simulations nor the Y03 simulations were evolved below redshifts of $z \\sim 20$ or had the spatial resolution to follow the chemo-thermal-hydrodynamics of cloud collapse on scales below $\\sim 0.1$ pc, making ``the fall of Pop III\" by radiative feedbacks alone quite uncertain and understudied. Indeed, Y03 found the Pop III global SFR was still rising at the end of their simulations, implying that we don't even know when the Pop III epoch peaked. Ricotti, Gnedin \\& Shull (2002) addressed this question using hydrodynamic cosmological simulations that included the feedback of both photo-dissociating and photo-ionizing radiation from primeval galaxies. Although their simulations were of even lower resolution than MBA and Y03 and parameterized star formation with ad hoc prescriptions, they found a cosmic star formation history quite similar to that predicted by Y03 despite small box size effects. They too could not simulate sufficiently large boxes to obtain converged results in the interesting redshift regime 20 > z > 6, but rather found that rare luminous objects strongly affected the average SFR within the simulation volume. It is fair to say that we have a much better idea when the first Pop III star formed in the universe than when the last one formed. Indeed, some models have been presented which predict Pop III stars forming as low as redshift 3 (Schneider et al. 2006). This review updates progress on direct numerical simulations of Pop III star formation by several groups since the reviews of Bromm and Larson (2004), Glover (2005), and Ciardi and Ferrara (2005). New high resolution simulations have been carried out by Yoshida et al. (2006), O'Shea \\& Norman (2007a) and Gao et al. (2007) which test the robustness of the basic picture described above by pushing the collapse to higher central densities, although not yet stellar densities, and by considering ensembles of simulations. The effects of a FUV background on Pop III star formation has recently been revisited by O'Shea \\& Norman (2007b) and by Wise and Abel (2007). It is found that \\h2 cooling remains important in halos with masses approaching $10^8 \\Msun$ and FUV backgrounds as strong as $J_{21}=1$. This suggests that the Pop III epoch may be more extended than previously thought and therefore occurring in rather different environments than originally simulated. Pop III star formation has been studied in new and different environments as well. Pop III star formation in warm dark matter models has been studied by O'Shea \\& Norman (2006) and by Gao \\& Theuns (2007). Finally, Pop III star formation in gas that has been pre-processed by earlier generation of Pop III stars has been simulated by O'Shea et al. (2005), Mesinger, Bryan \\& Haiman (2006), and by Yoshida et al. (2007a,b). The surprising result is that despite the diversity of the environments studied, it is found that Pop III.2 stars form in basically the same way as Pop III.1 stars and that they are massive, although they appear to be somewhat less massive than Pop III.1 based on lower cloud core temperatures and accretion rates. However, the new simulations show that environment and formation redshift/history can have a substantial effect on protostellar accretions rates, suggesting some spread in the primordial IMF. No 3D simulation has yet been done that follows the entire accretion history of the star, and therefore final masses are still uncertain. However, approximate integrations have been done assuming spherical symmetry which suggest that the entire protostellar envelope can be accreted (Omukai \\& Palla 2003, Yoshida et al. 2006, 2007a,b). Verifying this result with fully 3D simulations represents a grand challenge for the field. ", "conclusions": "" }, "0801/0801.0670_arXiv.txt": { "abstract": "{} {We conducted a search for brown dwarfs (BDs) and very low mass (VLM) stars in the 625~Myr-old Hyades cluster in order to derive the cluster's mass function across the stellar-substellar boundary.} {We performed a deep (I=23, z=22.5) photometric survey over 16~deg$^2$ around the cluster center, followed up with K-band photometry to measure the proper motion of candidate members, and optical and near-IR spectroscopy of probable BD and VLM members.} {We report the discovery of the first 2 brown dwarfs in the Hyades cluster. The 2 objects have a spectral type early-T and their optical and near-IR photometry as well as their proper motion are consistent with them being cluster members. According to models, their mass is 50 Jupiter masses at an age of 625~Myr. We also report the discovery of 3 new very low mass stellar members of the cluster, and confirm the membership of 16 others. We combine these results with a list of previously known cluster members to build the present-day mass function (PDMF) of the Hyades cluster from 50 Jupiter masses to 3~M$_\\odot$. We find the Hyades PDMF to be strongly deficient in very low mass objects and brown dwarfs compared to the IMF of younger open clusters such as the Pleiades. We interpret this deficiency as the result of dynamical evolution over the past few 100~Myr, i.e., the preferential evaporation of low mass cluster members due to weak gravitational encounters.} {We thus estimate that the Hyades cluster currently hosts about 10-15 brown dwarfs, while its initial substellar population may have amounted up to 150-200 members.} ", "introduction": "The determination of the initial mass function (IMF), i.e., the mass frequency distribution of objects formed in molecular clouds, yields strong constraints to star formation theories (Padoan \\& Nordlund 2002; Bate \\& Bonnell 2005; Larson 2005; Jappsen et al. 2005). In the last 10 years, the IMF has been derived from the most massive stars down to the substellar domain in a variety of environments : star forming regions (e.g. Luhman et al. 2003), young open clusters (e.g. Bouvier et al. 1998) and in the field (e.g. Reid et al. 1999). As brown dwarfs continuously cool down as they evolve (Chabrier et al. 2000), they are brighter when younger. Thus, the IMF could be derived down to a mass of a few Jupiter masses in star forming regions (Lucas et al. 2006; Caballero et al. 2007) and down to about 30 Jupiter masses in rich young open clusters (Moraux et al. 2003, 2007; Barrado et al. 2004; de Wit et al. 2006). While a large number of field brown dwarfs have also been discovered within a few tens of parsecs from the Sun, the derivation of the field IMF below 0.1~M$_\\odot$ still remains uncertain due to the lack of knowledge of the ages of substellar objects (Chabrier 2002; Allen et al. 2005; Cruz et al. 2007). In a previous series of paper, we derived the lower mass function of a number of young open clusters down to about 30~M$_{Jup}$ (see Moraux, Bouvier \\& Clarke 2005 and Bouvier, Moraux \\& Stauffer 2005 for short reviews) and discussed the implications of an apparently universal cluster mass function for star formation scenarios (Moraux et al. 2007). Here, we report an extended, deep optical survey of the Hyades cluster aimed at detecting substellar objects and deriving the cluster's mass function across the stellar-substellar boundary. The motivations for this survey are twofold. Firstly, all previous searches for substellar objects in the Hyades have failed to report any positive brown dwarf detections (Reid \\& Hawley 1999; Gizis, Reid \\& Monet 1999; Dobbie et al. 2002). The lowest mass members reported so far includes LH~0418+13, with a spectral type M8.5 and an estimated mass of 0.083~M$_\\odot$ (Reid \\& Hawley 1999), the $\\simeq$0.081~M$_\\odot$ companions of a couple of 0.1~M$_\\odot$ Hyades probable members, LP~415-20 and LP~475-855 (Siegler et al. 2003), and an unresolved very low mass companion (0.070-0.095~M$_\\odot$) in the spectroscopic binary RHy~403 (Reid \\& Mahoney 2000). Yet, the discovery of brown dwarfs at an age of 625~Myr (Perryman et al. 1998) would provide a unique benchmark to calibrate the substellar evolution models. Secondly, the Hyades cluster is dynamically evolved (Adams et al. 2002). A significant fraction of its initial low mass population is therefore expected to have drifted away beyond the cluster boundaries (e.g. Reid 1993). Deriving the lower mass function of such an evolved cluster would provide a direct measurement of the rate at which low mass cluster members evaporate and populate the field. In turn, this measurement would allow us to test the validity of N-body simulations of the dynamical evolution of young clusters (e.g. Kroupa 1995; Portegies Zwart et al. 2001). The Hyades cluster (Melotte 25, $\\alpha_{2000}$=04$^h$26$^m$54$^s$, $\\delta_{2000}$=+15$\\degr$52$\\arcmin$; $l$=180.05$\\degr$, $b$=-22.40$\\degr$) is the closest rich open cluster to the Sun. Perryman et al. (1998) derived its main structural and kinematical properties based on Hipparcos measurements~: a distance of 46.3$\\pm$0.27~pc, an age of 625$\\pm$50~Myr, a metallicity [Fe/H] of 0.14$\\pm$0.05, a present-day total mass of about 400~M$_\\odot$, a tidal radius of 10.3~pc, a core radius of 2.5-3.0~pc and negligible extinction on the line of sight. The large proper motion of the cluster ($\\mu\\simeq$100~mas yr$^{-1}$) can be easily measured from imaging surveys over a timeframe of only a few years, which helps in assessing cluster's membership. In Section~2, we describe the optical survey we performed over the central 16~deg$^2$ of the Hyades cluster, as well as follow-up K-band photometry and both optical and infrared spectroscopy. The proper motion of optically-selected candidate members is derived from optical and near-IR images obtained 2 to 3 years apart. In Section~3, we describe our selection of candidate members combining optical and inrafred photometry, as well as proper motion measurements and follow up spectroscopy. We thus report 21 probable members, of which 5 are new and 2 are the first Hyades brown dwarfs, with an estimated mass of $\\simeq$50~M$_{Jup}$. In Section~4, we discuss the spectral properties of the newly found low mass Hyades members, and derive a spectral type of T1 and T2 for the 2 brown dwarf candidates. We proceed in deriving the present-day mass function of the Hyades cluster from 0.050 to 3.0~M$_\\odot$, which we find to be strongly deficient in very low mass stars and brown dwarfs compared to the mass function of the younger Pleiades cluster. We discuss this result in the light of N-body simulations of the dynamical evolution of young open clusters. ", "conclusions": "From a deep, wide-field survey of the central region of the Hyades cluster, we identified new very low-mass cluster members, including 2 T-dwarfs with a mass of $\\simeq$ 50 Jupiter masses. The comparison of the Hyades lower mass function with that of younger clusters indicates that the cluster is strongly depleted in very low mass objects at an age of 625~Myr. We thus estimate a total of about 15 brown dwarfs in the present-day Hyades cluster compared to about 500 stars. This depletion appears to result from the preferential evaporation of the lowest mass cluster members over a timescale of several 100~Myr, as predicted by N-body models of the dynamical evolution of young clusters. A fraction of the substellar escapers should populate the solar neighborhood, and may be detected as members of the Hyades moving group from large scale surveys." }, "0801/0801.2169_arXiv.txt": { "abstract": "We study the extent and covering fraction of cool baryons around galaxies of different luminosity and mass, based on a survey of Mg\\,II $\\lambda\\lambda\\,2796, 2803$ absorption features near known galaxies. The initial sample consists of 13 galaxy and absorber pairs and 10 galaxies that do not produce Mg\\,II absorption lines to within sensitive upper limits. The redshifts of the galaxy and absorber pairs range from $z = 0.2067$ to 0.892 with a median of $z = 0.3818$. We find that galaxies at larger impact parameters produce on average weaker Mg\\,II absorbers. This anti-correlation is substantially improved when accounting for the intrinsic luminosities of individual galaxies. In addition, there exists a distinct boundary at $\\rho=R_{\\rm gas}$, beyond which no Mg\\,II absorbers are found. A maximum likelihood analysis shows that the observations are best described by an isothermal density profile and a scaling relation $R_{\\rm gas}=91\\times (L_B/L_{B_*})^{(0.35\\pm 0.05)}\\ h^{-1}$ kpc (or $69\\ h^{-1}$ kpc at $W(2796)=0.3$ \\AA) with a mean covering factor of $\\langle\\kappa\\rangle=80-86$ \\%. Together with the scaling relation between halo mass and galaxy luminosity inferred from halo occupation studies, this scaling of $R_{\\rm gas}$ indicates that gas radius is a fixed fraction of the dark matter halo radius. We compare our results with previous studies and discuss the implications of our analysis for constraining the baryon content of galactic halos and for discriminating between competing scenarios for understanding the nature of the extended gas. ", "introduction": "The Mg\\,II $\\lambda\\lambda\\,2796, 2803$ doublets are among the absorption features commonly seen in the spectra of distant quasars that are produced by intervening gaseous clouds along the quasar lines of sight. Their rest-frame absorption equivalent width $W(2796)$, ranging from $W(2796)\\apll 0.3$ \\AA\\ to $W(2796)>2$ \\AA\\ (e.g.\\ Steidel \\& Sargent 1990; Nestor \\etal\\ 2005; Prochter \\etal\\ 2006a), is found to represent the underlying gas kinematics (e.g.\\ Petitjean \\& Bergeron 1990; Churchill \\etal\\ 2000). Based on comparisons of the abundance ratios of various associated ions, these Mg\\,II absorption transitions are understood to arise primarily in photo-ionized gas of temperature $T\\sim 10^4$ K (Bergeron \\& Stas\\'inska 1986; Hamann 1997) and neutral hydrogen column density $N(\\hI)=10^{18}-10^{22}$ \\cmjj\\ (Churchill \\etal\\ 2000; Rao \\etal\\ 2006). In addition, surveys for galaxies near the observed Mg\\,II absorbers have often uncovered luminous galaxies at projected distances ($\\rho \\apll 50\\ h^{-1}$ kpc) and velocity separations ($\\Delta\\,v \\le 250$ \\kms) from the absorbers (Bergeron 1986; Lanzetta \\& Bowen 1990, 1992; Steidel \\etal\\ 1994; Zibetti \\etal\\ 2005; Nestor \\etal\\ 2007; Kacprzak \\etal\\ 2007). The relatively large associated $N(\\hI)$ and the small separation between Mg\\,II absorbers and galaxies along common lines of sight indicate that Mg\\,II absorbers may offer a sensitive probe of photo-ionized halo gas around galaxies at redshifts from $z=0.3$ to $z=2.3$ in the optical spectral window. Understanding the physical origin of these Mg\\,II absorbers bears significantly on all efforts to apply their known statistical properties for constraining the baryon content of dark matter halos on different mass scales (Tinker \\& Chen 2008). Various theoretical models have been developed in the past that describe the nature of the absorbers in the context of gas accretion, including ram pressure stripped gas from accreted satellites (Wang 1993), gravitationally bound cold gas in halo substructures (e.g., Sternberg \\etal\\ 2002), and condensed cold clouds in a thermally unstable hot halo (Mo \\& Miralda-Escude 1996; Maller \\& Bullock 2004; Chelouche \\etal\\ 2007). Insights to the origin of Mg\\,II absorbers can be obtained from their clustering amplitude on large scales ($\\apg\\,1\\,h^{-1}$ Mpc). The clustering of the absorbers is a consequence of dark matter halos in which they are found. Measurements of the galaxy and absorber cross--correlation function offer a means of quantifying the mean halo mass of the absorbers. High-mass halos are expected to be highly clustered, while low mass halos on average have weaker clustering strength. The cross--correlation function of Mg\\,II absorbers and luminous red galaxies (LRGs) have been measured at $\\langle z\\rangle=0.6$ by Bouch\\'e \\etal\\ (2006), who found that Mg\\,II absorbers of $W(2796)>0.3$ \\AA\\ cluster strongly with LRGs and that weaker absorbers with $W(2796)=0.3-1.15$ \\AA\\ on average arise in dark matter halos that are 10 times more massive than those of stronger absorbers with $W(2796)=2-2.85$ \\AA. Both the observed clustering amplitude and the inverse correlation between the mean halo mass and absorber strength are difficult to interpret. For example, the frequency distribution function of Mg\\,II absorbers show that there are on average 10 times more $W(2796)=1$ \\AA\\ absorbers than those of $W(2796)=2$ \\AA\\ (e.g.\\ Nestor \\etal\\ 2005; Prochter \\etal\\ 2006a) which, when combined with the clustering measurements, implies that the majority of absorbers reside in massive, highly biased halos. At the same time, we expect that observed Mg$^+$ ions originate primarily in photo-ionized gas of temperature $T\\sim 10^4$ K and the halo gas in massive dark matter halos becomes too hot for abundant Mg$^+$ to survive. In Tinker \\& Chen (2008; hereafter TC08), we developed a new technique that adopts the halo occupation framework for understanding the origin of QSO absorption-line systems. Specifically, the technique adopts a model density profile for the Mg$^+$ ions in individual dark matter halos. The ``cold'' baryon content of individual dark matter halos, as probed by the presence of Mg$^+$, is then constrained through matching the space density and clustering amplitude of dark matter halos with the observed frequency distribution function of Mg\\,II absorbers and their clustering amplitude. Our model allows the possibility that a predominant fraction of the gas in massive halos is shock heated to the virial temperature of the halo and becomes too hot to host abundant Mg\\,II absorbing clouds. Within the hot halo, we further allow the possibility that some dense, cold clouds may penetrate through as seen in high resolution cosmological simulations (Kravtsov 2003). The result of this halo occupation analysis is the probability function $P(W|M_h)$ that characterizes, for each dark matter halo of $M_h$, the total probability of finding a Mg\\,II absorber of equivalent width $W$. The halo occupation analysis shows that observations, including both the Mg\\,II absorber frequency distribution function and their clustering amplitude, demand a rapid transition in the halo gas content at $M_h^{crit}\\sim 10^{11.5}$ \\hmsol. Below $M_h^{crit}$, halos contain predominantly cold gas and therefore contribute significantly to the observed Mg\\,II statistics. Beyond $M_h^{crit}$, the cold gas fraction is substantially reduced and presumably the halo gas becomes too hot to maintain a large contribution to Mg\\,II absorbers. In order to reproduce the observed overall strong clustering of the absorbers and the inverse correlation between $W(2796)$ and halo mass $M_h$, roughly 5\\% of the gas in halos up to $10^{14}$ \\hmsol\\ is required to be cold. It is understood under our model that {\\it the $W(2796)$ vs.\\ $M_h$ inverse correlation arises mainly as a result of an elevated clustering amplitude of $W(2796)\\apll 1$ \\AA\\ absorbers, rather than a suppressed clustering strength of $W(2796)\\apg 2$ \\AA\\ absorbers.} The clustering amplitude of weak absorbers is elevated because of the presence of cold streams in massive hot halos. The amount of cold gas in clusters is small, therefore these halos can only contribute to weak absorbers. The initial results of our halo occupation analysis demonstrate that combining known statistics of dark matter halos with a simple model for their gas content can already reproduce the statistical properties known for Mg\\,II absorbers. It provides a simple prescription for populating baryons in dark matter halos that can be compared directly to results from numerical simulations and offer insights for understanding the physics of gas accretion in halos of all mass scales. However, some uncertainties remain in the halo occupation analysis. First, a key parameter that constrains the distribution of Mg$^+$ ions in individual dark matter halos is their incidence rate, $\\kappa(M_h)$, per halo. It specifies the total probability of detecting an absorber in a halo of mass $M_h$. In the halo occupation analysis of TC08, we find that dark matter halos of $M_h=10^{11.5-12.5}$ \\hmsol\\ on average have 100\\% incidence rate of Mg$^+$ ions, $\\langle\\kappa(M_h)\\rangle=1$, at $R_{\\rm gas}\\le R_{\\rm 200}/3$ for Mg$^+$ ions and $\\langle\\kappa(M_h)\\rangle=0$ at larger radii.\\footnote{The halo size $R_{\\rm 200}$ corresponds to the radius, within which the enclosed mean density is $200\\times\\bar{\\rho}_m$ and $\\bar{\\rho}_m$ is the background density. It is motivated by estimates of the virial radius both from the spherical collapse model and from $N$-body simulations that predict $\\approx 180\\times\\bar{\\rho}_m$.} In halos of lower masses, $\\langle\\kappa(M_h)\\rangle$ declines sharply. While a 100\\% incidence rate shows that every dark matter halo hosts a uniform gaseous halo of Mg$^+$ ions with size $R_{\\rm gas}$, the low $\\langle\\kappa(M_h)\\rangle$ is more difficult to interpret because $\\langle\\kappa(M_h)\\rangle$ represents a mean value averaged over all halos of mass $M_h$. A low $\\langle\\kappa(M_h)\\rangle$ may be a result of only a small fraction of dark matter halos containing extended distributions of Mg$^+$ ions or a result of a small covering factor of Mg$^+$ in all halos. In addition, we have assumed in our initial analysis that the gaseous extent is related to halo mass according to $R_{\\rm gas}= R_{{\\rm gas}*}\\,[M_h/(10^{12}\\,\\hmsol)]^{\\beta}$, and $R_{{\\rm gas}*}=50\\ h^{-1}$ physical kpc and $\\beta=1/3$. This scaling relation follows the theoretical expectation between virial radius and halo mass, but a higher $R_{\\rm gas}$ would naturally lower $\\kappa$ for a fixed frequency distribution function of Mg\\,II absorbers. Second, the physical origin of the cold clouds probed by the Mg\\,II absorption transitions is ambiguous. While the mass scale at which halo gas is found to experience a transition from a cold-mode dominated state to a hot-mode dominated state agrees well with the expectation of theoretical models established to characterize the gas accretion history in dark matter halos (e.g.\\ Birnboim \\& Dekel 2003; Kere{\\^s} \\etal\\ 2005; Dekel \\& Birnboim 2006; Birnboim \\etal\\ 2007), the observed cold-hot transition can also be interpreted as a declining formation efficiency of cool clouds in a hot halo with increasing halo mass (e.g.\\ Mo \\& Miralda-Escude 1996; Maller \\& Bullock 2004). Both of these scenarios predict that halo gas is primarily cold at $M_h\\apll 10^{11.5}\\,\\hmsol$. On the other hand, Bouch\\'e \\etal\\ (2006) have argued that the inverse correlation between $W(2796)$ and clustering amplitude is suggestive of the absorbers arising in non-virialized gas flows, such as starburst winds. While dense clumps in starburst driven outflows are expected to contribute to some fraction of the observed Mg\\,II absorbers, an important quantity to specify is the significance of this fraction. If a large fraction of Mg\\,II absorbers originate in cold dense clumps in starburst winds, then one must also account for the more complex star formation physics in efforts to apply known Mg\\,II statistics for constraining the gas content of dark matter halos. We defer to \\S\\ 5.3 for a more detailed discussion of the caveats in interpreting the known properties of Mg\\,II absorbers as evidence to support their origin in starburst winds. Direct constraints on the covering fraction of Mg\\,II absorbing gas versus halo mass not only help to break the degeneracy between the fraction of dark matter halos containing cold baryons and the extent of cold gas in individual halos, in principle they also serve to discriminate between the starburst wind scenario and gas accretion. Under the starburst wind scenario, superwind driven outflows are expected to proceed over a finite angular span along the minor axis (e.g.\\ Heckman \\etal\\ 1990; Veilleux \\etal\\ 2005), and result in only partial covering of the halos. In addition, because at a given epoch only some fraction of galaxies are found at starburst or post-starburst stages, we would expect a fraction of dark matter halos to contain extended Mg\\,II absorbing gas due to outflows. To obtain empirical constraints on the extent and covering fraction of Mg$^+$ ions around galaxies of different luminosity and mass, we have initiated a survey of Mg\\,II absorbers around known galaxies at small projected distances ($\\rho\\apll 100\\ h^{-1}$ kpc) to a background QSO. Accounting the presence or absence of Mg\\,II absorbers for an unbiased sample of galaxies at different impact parameters yields a statistical estimate of the gas covering fraction around field galaxies. A detailed comparison between the absorber strength and galaxy properties further constrains the density profile and extent of the gas. The primary objectives of our study are (1) to improve the uncertainties of our halo occupation analysis by obtaining empirical measurements of $R_{\\rm gas}$ and $\\beta$, and (2) to examine whether starburst driven outflows are the predominant mechanism for producing the observed Mg\\,II absorbers by obtaining an empirical constraint of the gas covering factor $\\kappa$. Here we present the initial results of our study. This paper is organized as follows. In Section 2, we describe the design of our experiment for constraining the extent and covering fraction of Mg$^+$ ions. In Section 3, we describe the imaging and spectroscopic data available for our experiment. In Section 4, we examine the correlation between Mg\\,II absorption strength and galaxy properties, such as the impact parameter and luminosity. We compare our results with previous studies and discuss the implications of our analysis in Section 5. We adopt a $\\Lambda$CDM cosmology, $\\Omega_{\\rm M}=0.3$ and $\\Omega_\\Lambda = 0.7$, with a dimensionless Hubble constant $h = H_0/(100 \\ {\\rm km} \\ {\\rm s}^{-1}\\ {\\rm Mpc}^{-1})$ throughout the paper. ", "conclusions": "Using the sample of 13 galaxy--Mg\\,II absorber pairs and 10 galaxies at $\\rho\\apll 100\\ h^{-1}$ kpc that do not give rise to Mg\\,II absorption to a sensitive upper limit, we find that galaxies at larger impact parameters produce on average weaker Mg\\,II absorbers. This anti-correlation is substantially improved, when accounting for the intrinsic luminosity of individual galaxies. In addition, there exists a distinct boundary at $\\rho=R_{\\rm gas}$, beyond which no Mg\\,II absorbers with $W(2796)\\ge 0.01$ \\AA\\ are found. A maximum likelihood analysis shows that the observations are best described by an isothermal density profile of the gas and a scaling relation $R_{\\rm gas}=91_{-8}^{+3}\\times (L_B/L_{B_*})^{(0.35\\pm 0.05)}\\ h^{-1}$ kpc with a mean covering factor of $\\langle\\kappa\\rangle=80-86$ \\% at $\\rho\\le R_{\\rm gas}$. The best-fit profile applies to galaxies of $0.03\\,L_{B_*}$ to $1.6\\,L_{B_*}$ and therefore predominantly cold-mode halos. For higher mass halos, we expect from TC08 that the corresponding absorption strength for a given $\\rho$ is reduced, namely a smaller $\\bar{W}_0$ in Equation (1). In this section, we discuss the implications drawn from the results of our analysis. \\subsection{Comparisons with Previous Studies} Adopting the best-fit scaling relation and the gaseous profile described in Equation (1), we derive a corresponding halo size of $\\rho\\approx 69\\ h^{-1}$ physical kpc for galaxies of $M_B-5\\,log\\,h=-19.8$ at $W(2796)=0.3$ \\AA. This halo size is consistent with the finding of Lanzetta \\& Bowen (1990) and the expectation inferred from known Mg\\,II frequency function (Nestor \\etal\\ 2005; see the descriptions in \\S\\ 5.2), but is larger than what was reported in Steidel (1995). Steidel (1995) reported a characteristic size of Mg\\,II halo of $R_*\\approx 35\\ h^{-1}$ kpc for $L_*$ (corresponding to $M_{B_*}-5\\,\\log\\,h=-19.7$ for the cosmology adopted in this paper) galaxies, but it is not clear at what $W(2796)$ limit the halo size was estimated. At $\\rho=35\\ h^{-1}$ kpc, we expect $\\bar{W}(2794)=0.9$ \\AA\\ from the best-fit model. The high gas covering factor derived from our analysis is consistent with what is inferred from the sample of Kacprzak \\etal\\ (2008). Applying the best-fit luminosity scaling relation of gaseous extent, we find that four of the 28 galaxies within the expected gaseous extent for $W_r(2796)> 0.3$ \\AA\\ absorbers in the Kacprzak \\etal\\ sample have a corresponding Mg\\,II weaker than the 0.3 \\AA\\ threshold. The inferred covering fraction is $\\approx 86$\\%. At the same time, this high covering fraction disagrees with the value reported in Tripp \\& Bowen (2005), the only other on-going survey program designed to constrain gas covering fraction by searching for corresponding Mg\\,II absorbers at the locations of known {\\it field} galaxies\\footnote{We note two earlier surveys by Bechtold \\& Ellingson (1992) and Bowen \\etal\\ (1995) that include a dominant fraction of galaxies in dense cluster environment. Both studies reported a substantially lower covering fraction, $\\langle\\kappa\\rangle\\apll 20$ \\% for $W(2796)>0.1$ absorbers. The low covering fraction of strong absorbers is understood, because galaxies clusters reside in massive dark matter halos of $M_h> 10^{14}\\,\\hmsol$ and the fraction of ``cold'' gas in these massive halos is expected to be reduced (see TC08).}. The authors obtained follow-up spectra of background QSOs that are located within $\\rho=10-39\\ h^{-1}$ physical kpc of 20 known galaxies at $z_{\\rm gal}=0.31-0.55$. They found that 10 of the 20 galaxies do not produce Mg\\,II absorbers to within a 2-$\\sigma$ upper limit of $W(2796)=0.1$ \\AA, and concluded that the incidence of Mg$^+$ ions is $\\approx\\,50$ \\% at $\\rho\\apll 40\\ h^{-1}$ kpc in individual galactic halos. Because more luminous galaxies have larger extended Mg\\,II halos according to the scaling relation presented in Equations (2), (6), and (7), this discrepancy can be understood if many of the galaxies in Tripp \\& Bowen are located at impact parameters that are larger than their expected $R_{\\rm gas}$. Recall that we obtained a similar estimate of $\\langle\\kappa\\rangle\\approx 57$ \\% at $\\rho< 100\\ h^{-1}$ kpc (or $\\langle\\kappa\\rangle\\approx 50$ \\% at $\\rho< 45\\ h^{-1}$ kpc), {\\it prior to} accounting for the scaling relation between $R_{\\rm gas}$ and $L_B$ in individual galaxies (\\S\\ 4). In addition, the sample in Tripp \\& Bowen includes galaxies of luminosity ranging from $0.3\\,L_*$ to $5\\,L_*$. Those galaxies that are more luminous than $L_*$ are also expected to reside in more massive halos. At $z=0$, $5\\,L_*$ galaxies are expected to reside in $> 10^{13}\\,\\hmsol$ halos (Tinker \\etal\\ 2007). If the host dark matter halos are substantially more massive than $10^{12.0}\\,\\hmsol$, then we expect from the halo occupation analysis in TC08 that they do not contribute significantly to strong Mg\\,II absorbers. In summary, further inspections of the properties of these galaxies and estimates of the gas covering fraction $\\kappa$ as a function of radius and $W(2796)$ are crucial for resolving this discrepancy. \\subsection{The Origin of Mg\\,II Absorbers} The scaling relation and the constraint of $\\langle\\kappa\\rangle$ together also allow us to test whether the Mg\\,II absorbers trace a significant and representative portion of the galaxy population or merely a special class such as those galaxies undergoing a starburst episode. If extended Mg\\,II halos are a generic feature of field\\footnote{We define field galaxies as those that are not associated with a known cluster.} galaxies at all epochs, then we can derive the expected number density of Mg\\,II absorbers by combining known space density of galaxies (versus luminosity) and the product of the cross section and covering factor of the Mg$^+$ ions. An agreement with observational results from surveys of Mg\\,II absorbers would challenge the scenario that attributes a large fraction of these absorbers to starburst systems\\footnote{However, we will not to able to rule out a scenario in which the blow-out winds are common in most galaxy histories. Given enough time, long-lived outflowing gas gets mixed in with existing halo gas and newly accreted material from the IGM. There is little distinction at this point between outflows and accreted clouds. It is also not clear whether the outflow material would maintain the same degree of complex gas kinematics as those that are closer to the parent starburst regions.}. The predicted number density of Mg\\,II absorption systems arising in the extended gaseous halos of galaxies may be evaluated according to \\begin{eqnarray} \\frac{d\\,{\\cal N}(W\\ge W_{lim})}{d\\,z} & = & \\frac{c}{H_0} \\frac{(1 + z)^2}{\\sqrt{\\Omega_{\\rm M}\\,(1+z)^3 + \\Omega_{\\rm \\Lambda}}} \\\\ & & \\times \\int_0^\\infty d\\left(\\frac{L_B}{L_{B_*}}\\right)\\,\\Phi(L_B,z)\\,\\sigma(L_B,W_{lim})\\,\\kappa, \\nonumber \\end{eqnarray} where $c$ is the speed of light, $\\Phi(L_B,z)$ is the galaxy luminosity function, $\\sigma$ is the gas cross section for producing Mg\\,II absorbers of $W(2796)\\ge W_{lim}$ that scales with galaxy $B$-band luminosity, and $\\kappa$ is the halo covering factor. Substituting the scaling relationship according to Equations (2), (6), and (7), we find \\begin{eqnarray} \\frac{d\\,{\\cal N}(W\\ge W_{lim})}{d\\,z} & = & \\frac{c}{H_0}\\frac{\\pi\\,\\langle\\kappa\\rangle\\,[R'_{{\\rm gas}*}(W_{lim})]^2\\,(1 + z)^2}{\\sqrt{\\Omega_{\\rm M}\\,(1+z)^3 + \\Omega_{\\rm \\Lambda}}} \\\\ & & \\times\\int_0^\\infty d\\left(\\frac{L_B}{L_{B_*}}\\right)\\,\\left(\\frac{L_B}{L_{B_*}}\\right)^{2\\,\\beta}\\,\\Phi(L_B,z)\\,, \\nonumber \\end{eqnarray} where $R'_{{\\rm gas}*}(W_{lim})$ is the extent of Mg\\,II halos that corresponds to $W(2796)=W_{lim}$. To evaluate Equation (9), we adopt the luminosity function of the blue galaxy population at $z_{\\rm gal}=0.3-0.5$ from Faber \\etal\\ (2007), which is characterized by $M_{B_*}-5\\,\\log=-19.8$, $\\phi_*=9.6\\times 10^{-3}\\ h^{3}$ Mpc$^{-3}$, and a faint-end slope $\\alpha=-1.3$. Including galaxies of $L_B\\ge 0.01\\,L_{B_*}$ and our best-fit $R_{{\\rm gas}*}$ and $\\langle\\kappa\\rangle$, Equation (9) yields $d\\,{\\cal N}/dz=0.028-0.03$ for $W(2796)\\ge 2$ \\AA\\ absorbers, $d\\,{\\cal N}/dz=0.21-0.23$ for $W(2796)\\ge 1$ \\AA\\ absorbers, and $d\\,{\\cal N}/dz=0.88-1.07$ for $W(2796)\\ge 0.3$ \\AA\\ absorbers at $z=0.5$. The predictions based on galaxies more luminous than $0.01\\,L_*$ agree well with the observed $d\\,{\\cal N}/d\\,z$ from Nestor \\etal\\ (2005), lending strong support for the hypothesis that the Mg\\,II absorbers trace a significant and representative portion, rather than a sub-class, of the galaxy population. The model profile of Equation (1) has been proved to be a good representation of the galaxy and absorber pairs at $\\rho=16.3-100\\ h^{-1}$ kpc, when the scaling relation of Equations (2), (6), and (7) is accounted for. An important test is to examine whether the model remains a good fit at smaller $\\rho$ and strong absorbers. Identifying galaxy--absorber pairs at small impact parameters is difficult, because the glare of the background QSO often prohibits identification of faint galaxies at angular distances $\\theta<2''$ that corresponds to $\\sim 10\\ h^{-1}$ physical kpc at $z\\apll 0.5$. Recent studies of intervening absorption-line systems along the lines of sight toward the optical afterglows of distant $\\gamma$-ray bursts (GRB) have uncovered a number of strong Mg\\,II absorbers (e.g.\\ Prochter \\etal\\ 2006b). Unlike QSOs, GRB afterglows disappear after a while, permitting exhaustive searches of faint galaxies at $\\theta<2''$ that give rise to strong Mg\\,II absorbers along the lines of sight. A redshift survey of galaxies along the sightline toward GRB\\,060418 ($z_{\\rm GRB}=1.491$) has uncovered the absorbing galaxies for three strong Mg\\,II absorbers at $z=0.603-1.107$ (Pollack \\etal\\ 2008, in preparation), similar to the redshift range covered by our sample. The rest-frame $B$-band magnitudes of the galaxies range from $M_B-5\\,\\log\\,h=-16.8$ to $M_B-5\\,\\log\\,h=-18.3$; Impact parameters of the galaxies range from $\\rho=7.5\\ h^{-1}$ kpc to $\\rho=16.5\\ h^{-1}$ kpc. Adopting the best-fit scaling relation, we include the three galaxy--Mg\\,II absorber pairs in the $W(2796)$ versus $\\rho$ plot for comparison (star points in Figure 5). The good agreement between the best-fit model and the absorbers found along GRB sightlines provides further support for our model at small impact parameters. \\begin{figure} \\includegraphics[scale=0.45]{f5.eps} \\caption{The $W(2796)$ versus $\\rho$ correlation scaled by galaxy $B$-band luminosity presented in Figure 4, but including pairs from Pollack \\etal\\ (2008; star points) and Bouch\\'e \\etal\\ (2007; points with horizontal errorbars) for comparisons. The solid curves represent the best-fit model based on an isothermal density profile, with $a_h=0.1\\,R_{\\rm gas}$ (top curve) and $a_h=0.2\\,R_{\\rm gas}$ (bottom curve). The pairs of Pollack \\etal\\ are for galaxies and Mg\\,II absorbers identified at $z=0.6-1$ along the sightline toward GRB\\,060418. The errorbars in these data points are smaller than the size of the symbols. The pairs of Bouch\\'e \\etal\\ are galaxies found at the positions of known Mg\\,II absorbers based on the presence of H$\\alpha$ emission lines. Errorbars represent the uncertainties of the inferred $M_B$ from the $M_B$ versus $L(\\ha)$ correlation published in Tresse \\etal\\ (2002). Crosses represent the absorbers in the Bouch\\'e sample that are likely ($\\apg 40$ \\% probability) to be a damped \\lya\\ absorber of $N(\\hI)\\ge 2\\times 10^{20}$ \\cmjj\\ based on the presence of strong Fe\\,II and Mg\\,I transitions (Rao \\etal\\ 2006).} \\end{figure} \\subsection{Are Strong [$W(2796)>1$ \\AA] Absorbers at $z\\sim 1$ Produced Primarily in Starburst Outflows?} As described in \\S\\ 1, the large line width of $W(2796)>1$ \\AA\\ Mg\\,II absorbers has recently been interpreted as due to outflow motions in starburst driven winds. This scenario was initially motivated by the observed inverse correlation between $W(2796)$ and the clustering amplitude of Mg\\,II absorbers at $\\langle z\\rangle=0.6$ (Bouch\\'e \\etal\\ 2006). In TC08, we have shown that this inverse correlation can be understood as due to an elevated clustering amplitude of $W(2796)\\apll 1$ \\AA\\ absorbers from contributions of survived cold clouds in massive hot halos, rather than a suppressed clustering strength of $W(2796)\\apg 2$ \\AA\\ absorbers. In addition, we have shown in \\S\\S\\ 4, 5.1, and 5.2 that extended Mg\\,II halos are a generic feature of field galaxies over a wide luminosity range. The predicted number density of the absorbers from adopting the scaling relation, Equations (2), (6), and (7), and the galaxy luminosity function agrees well with observations. Specifically, we expect $d\\,{\\cal N}/dz\\approx 0.03$ for $W(2796)\\ge 2$ \\AA\\ absorbers, in comparison to the observed $d\\,{\\cal N}/dz=0.03-0.04$ for $W(2796)\\ge 2$ \\AA\\ absorbers at $z\\approx 0.5$ from Nestor \\etal\\ (2005). While there is no clear indicator one can apply to unambiguously distinguish between gas inflows and outflows in the distant universe, we discuss several caveats related to the starburst outflow scenario. If the absorbers are produced in starburst driven outflows, then a large fraction of the absorbing galaxy population might be expected to exhibit disturbed morphology (Mobasher \\etal\\ 2004). The Mg\\,II absorbing galaxies in our sample exhibit regular disk morphologies in the HST images shown in Figure 2. In particular, the strong absorber found with $W(2796)=1.55$ \\AA\\ at $z=0.892$ toward 3C336 is associated with an edge-on disk galaxy at $\\rho= 16.3\\ h^{-1}$ kpc. The three Mg\\,II absorbers of $W(2796)>1$ \\AA\\ found toward GRB\\,060418 also exhibit regular disk structures in high-resolution images obtained both in space using HST and on the ground using LGAO on the Keck telescopes (Pollack \\etal\\ 2008, in preparation). Kacprzak \\etal\\ (2007) also showed that there is little correlation between $W_r(2796)$ and galaxy asymmetry for Mg\\,II absorbers of $W_r(2796)>1.4$ \\AA. While some strong Mg\\,II absorbers have been presented to show evidence of arising in expanding superbubble shells (Bond \\etal\\ 2001), available high-resolution images of other strong Mg\\,II absorbers do not support outflow being principally responsible for the observed strong absorbers. This conclusion is also consistent with the spectral properties of Mg\\,II-selected galaxies published by Guillemin \\& Bergeron (1997) On the other hand, Bouch\\'e \\etal\\ (2007) presented new observations from a targeted search for corresponding H$\\alpha$ emission in the vicinity of known Mg\\,II absorbers. These authors successfully identified H$\\alpha$ emission near 14 of 21 Mg\\,II absorbers with $W(2796)>1.9$ \\AA\\ at $z\\sim 1$. The integrated raw\\footnote{No extinction correction is applied to $L(\\ha)$ in our analysis, because the dust content is expected to vary substantially in $z\\sim 1$ galaxies (e.g.\\ Tresse \\etal\\ 2002) and is unknown for galaxies in the Bouch\\'e \\etal\\ sample. Bouch\\'e \\etal\\ (2007) applied a constant correction to all observed \\ha\\ flux in their calculation, resulting in a simple scaling of the observed values. The scatter in their measurements induced by dust remains the same. For consistency, we consider in the following discussion only observed values without extinction correction both for the absorbing galaxies and for comparison results, such as the \\ha\\ luminosity function and the $M_B-L(\\ha)$ correlation, from the literature.} \\ha\\ luminosities range from $L_{\\ha}=6.8\\times 10^{40}\\ h^{-2}$ erg s$^{-1}$ to $L_{\\ha}=8.4\\times 10^{41}\\ h^{-2}$ erg s$^{-1}$. The impact parameters range from $\\rho=1.4\\ h^{-1}$ kpc to $\\rho=35\\ h^{-1}$ kpc. Adopting a mean extinction correction $A_V=0.8$ mag, the authors inferred a mean star formation rate of $\\langle{\\rm SFR}\\rangle=5.9\\ h^{-2}$ M$_\\odot$ yr$^{-1}$ for these galaxies. The \\ha\\ luminosity function at $z\\sim 1$ has been studied by a number of groups (see Doherty \\etal\\ 2006 for a list of references). It is characterized by a Schechter function of $L_{{\\ha}_*}=5\\times 10^{41}\\ h^{2}$ erg s$^{-1}$, $\\phi_*=0.013\\ h^3\\,{\\rm Mpc}^{-3} $, and $\\alpha=-1.3$. The galaxies detected in \\ha\\ by Bouch\\'e \\etal\\ correspond to $0.14-1.68\\,L_{\\ha_*}$ at $z\\sim 1$, and therefore have only modest $L(\\ha)$. Integrating the \\ha\\ luminosity function, we obtain a mean space density of 0.034 $h^{3}$ Mpc$^{-3}$ for galaxies of brighter than $0.1\\,L_{{\\ha}_*}$. This is already 50\\% of all galaxies more luminous than $0.01\\,L_{B_*}$ at $z\\sim 1$, according to the best-fit luminosity function of the blue galaxy population from Faber \\etal\\ (2007). This exercise shows that \\ha\\ emission is commonly seen in $z\\sim 1$ galaxies and the presence of modest $L(\\ha)$ does not argue for the presence of starburst outflows, unless the outflow material is long-lived. For comparison, we include the 14 galaxy and absorber pairs from Bouch\\'e \\etal\\ (2007) in the $W(2796)$ versus $\\rho$ plot in Figure 5, applying the same scaling relation described in Equations (2), (6), and (7). We estimate $M_B$ of these galaxies based on the published $M_B-L(\\ha)$ correlation in Tresse \\etal\\ (2002). We adopt the observed $L(\\ha)$ and the scaling relation $M_B=73.5-2.27\\times\\log\\,L_(\\ha)$ that best describes the data in Tresse \\etal\\ (2002). At small impact parameters, the model predictions depend sensitively on the choice of $a_h$ in Equation (1). In addition to the fiducial model, we also include in the figure a model with $a_h=0.1\\,R_{\\rm gas}$. Nine of the 14 Mg\\,II absorbers are considered likely ($>40$ \\%) damped \\lya\\ absorbers (DLAs) of $N(\\hI)\\ge 2\\times 10^{20}$ \\cmjj\\ based on the presence of strong Fe\\,II and Mg\\,I transitions (Rao \\etal\\ 2006). DLAs are the high-redshift analog of neutral gas regions that resembles the disks of nearby luminous galaxies (e.g.\\ Wolfe \\etal\\ 2005). We therefore expect that the impact parameter distribution of galaxy-DLA pairs represents the typical extent of H\\,I disks at high redshifts. We highlight the potential DLAs in the Bouch\\'e \\etal\\ sample as crosses in Figure 5. After applying the best-fit scaling relation for Mg\\,II absorbers, we find that the impact parameters of these galaxy and potential DLA pairs range from $8\\ h^{-1}$ kpc to as high as $38\\ h^{-1}$ kpc. The three largest separation pairs have $\\rho'>28\\ h^{-1}$ kpc, where $\\rho'$ is the luminosity scaled projected distance. The scaling of impact parameter is justified, because smaller/fainter galaxies have on average smaller \\hI\\ disks (e.g.\\ Cayatte \\etal\\ 1994). The large impact parameters, together with the likely high $N(\\hI)$, suggest that the galaxies responsible for these absorbers (presumably at smaller angular distances from the QSOs) may still be missing. For the remaining sample, our model is in reasonably good agreement with their data. In summary, we find a lack of empirical measurements that argue for starburst outflows to be a dominant mechanism for producing the observed strong Mg\\,II absorbers at $z\\sim 1$. On the other hand, both the observed number density of Mg\\,II absorbers and the $W(2796)$ versus $\\rho$ correlation can be explained by extended Mg\\,II halos around typical field galaxies. The fraction of Mg\\,II absorbers originating in starburst outflows is therefore expected to be small at $z\\sim 1$, unless starburst outflows are a common feature in field galaxies. We caution, however, that the same conclusion cannot be automatically applied to absorbers at $z\\sim 2$ before a similar analysis is performed. \\subsection{Constraints on the Extended Gaseous Halos Around Galaxies} The sharp decline of $W(2976)$ found at $\\rho\\approx 90\\ h^{-1}$ kpc in Figure 4 clearly indicates that Mg$^+$ ions have a finite extent in the extended gaseous halos around galaxies. This is reminiscent of the sharp boundary between the C\\,IV absorbing and non-absorbing regions around $\\langle z\\rangle=0.4$ galaxies reported in Chen \\etal\\ (2001b). Given the nature of a photo-ionized gas (Bergeron \\& Stas\\'inska 1986; Hamann 1997), we examine whether the finite extent of Mg$^+$ ions can be understood as due to photo-ionization of the absorbing clouds at large radii. In a two-phase halo model (e.g.\\ Mo \\& Miralda-Escud\\'e 1996), cold clouds that are responsible for producing the absorption features are pressure confined in a hot medium. As the total gas density declines toward larger radii, the density of the cold clouds also decreases and the clouds become optically thin to the ultraviolet photons that can quickly ionize the Mg$^+$ ions (I.P.\\,$=$\\,15.035 eV). With a known background radiation field, the location where the ionization transition occurs may be adopted to constrain the density of cold gas at large radii. We carry out the photo-ionization analysis, using the Cloudy software (Ferland \\etal\\ 1998; version 06.02), and calculate the ionization fractions of Mg$^+$ ions for a grid of models with different total gas density and metallicity. We adopt the spectral shape of the UV background radiation field of Haardt \\& Madau in Cloudy for the incident ionizing flux, and scale the radiation intensity to the background UV radiation field inferred from QSO proximity effect (Scott \\etal\\ 2002). We adopt a plane parallel geometry for the clouds with a thickness of 100 pc, which is motivated by the observed coherent length of Mg\\,II absorbers toward lensed QSOs (Rauch \\etal\\ 2002). The slab of gas is illuminated by the ionization radiation field on both sides. \\begin{figure} \\begin{center} \\includegraphics[scale=0.4]{f6.eps} \\caption{Ionization fractions of Mg$^+$ ions and H$^0$ atoms versus total gas density $n_{\\rm H}$ for photo-ionized gas. The values are calculated using the Cloudy software (Ferland \\etal\\ 1998; version 06.02) for gaseous clouds of plane parallel geometry and 100 pc thickness. The input ionizing radiation intensity is set to be $J_\\nu(912\\AA)=7\\times 10^{-23}$ erg s$^{-1}$ Hz$^{-1}$ cm$^{-2}$ Sr$^{-1}$ suitable for the $z<1$ universe (Scott \\etal\\ 2002). We perform the calculations, assuming solar (dotted curve for Mg and dot-dashed curve for H) and $1/10$ solar metallicity (solid curve for Mg and dashed curve for H).} \\end{center} \\end{figure} Figure 6 shows the expected fractions of Mg$^+$ ions (solid and dotted curves) as a function of the total gas density $n_{\\rm H}$. We include the calculated fraction of H$^0$ atoms (dashed and dash-dotted curves) for comparison. The solid and dashed curves are for gas of $1/10$ solar metallicity, while the dotted and dash-dotted curves are for gas of solar metallicity. Our photo-ionization calculations indicate that indeed the fraction of Mg$^+$ declines linearly at $n_{\\rm H}\\apll 0.025\\ {\\rm cm}^{-3}$ for a mean radiation intensity of $J_\\nu(912\\AA)=7\\times 10^{-23}$ erg s$^{-1}$ Hz$^{-1}$ cm$^{-2}$ Sr$^{-1}$ (which is appropriate for the observed background radiation field at $z<1$ Scott \\etal\\ 2002). However, the transition is not as sharp as what is seen for hydrogen atoms. Including the estimated ionization fraction from the Cloudy models and adopting the density profile parameterized in Equation (1), we derive the expected correlation between $W_r(2796)$ and $\\rho$ and find that photo-ionization is insufficient for explaining the presence of the sharp decline in the observed Mg\\,II absorption strength. This extent of Mg\\,II gaseous halos ($\\approx 90\\ h^{-1}$ kpc around typical $L_*$ galaxies) is remarkably similar to the extent of C\\,IV gaseous halo ($\\approx 110\\ h^{-1}$ kpc around typical $L_*$ galaxies, after correcting for the $\\Lambda$-cosmology) reported by Chen \\etal\\ (2001b). On the basis of 14 galaxy and C\\,IV absorber pairs and 36 galaxies that do not produce corresponding C\\,IV absorption lines to within sensitive upper limits, these authors find that the extent of C\\,IV-absorbing gas around galaxies scales with galaxy $B$-band luminosity as $R_{\\rm gas}({\\rm C\\,IV}) \\propto 110 \\times L_B^{0.5 \\pm 0.1}\\ h^{-1}$ kpc (corrected for the $\\Lambda$ cosmology). In addition, there exists a sharp boundary between C\\,IV absorbing and non-absorbing regions at $R_{\\rm gas}({\\rm C\\,IV})$. But the scatter in the observed C\\,IV absorption strengths at $\\rho< R_{\\rm gas}({\\rm C\\,IV})$ between different galaxies appears to be significantly larger than what is seen in Figure 4, and there is a lack of anti-correlation between the observed C\\,VI absorbing strength and galaxy impact parameter. Given the distinct ionization potentials of the C$^{3+}$ and Mg$^+$ ions, {\\it the location where the abrupt transition between absorbing and non-absorbing regions occurs can be understood as a critical radius below which cool clouds can form and stablize in an otherwise hot halo}. This two-phase model to interpret QSO absorption line systems was formulated in Mo \\& Miralda-Escud\\'e (1996) and later re-visited by Maller \\& Bullock (2004). Taking into account the appropriate photo-ionization condition of the cold clouds, Mo \\& Miralda-Escud\\'e presented the expected Mg\\,II and C\\,IV absorption strength versus radius for halos of different mass. Their models showed a lack of radial dependence on the C\\,IV absorption strength and a tight correlation between Mg\\,II $W_r(2796)$ and $\\rho$ for halos of a wide range of mass. The differences in the radial dependence of different ions is understood as being due to the photo-ionization condition of cold clouds in the UV background radiation field. These model expectations agree with observational findings very well. The agreement between observations and a simple two-phase model is encouraging, although it is not clear how the two-phase model applies to cold-mode halos (e.g.\\ Dekel \\& Birnboim 2006). Measurements of the relative abundances between C$^{3+}$ and Mg$^+$ as a function of radius will provide additional support for the origin of these metal-line absorbers, further constraining the metallicity and density of halo gas around galaxies. \\subsection{Implications for the Halo Occupation Analysis and Future Work} Our initial halo occupation analysis (TC08) included two principal assumptions for the extent of Mg$^+$. First, we assumed that the extent of the cold gaseous halo is $R_{\\rm gas,12}=50\\,h^{-1}$ physical kpc for $M_h=10^{12}\\,\\hmsol$ halos, which is roughly $1/3$ of the halo radius $R_{200}$ at $z=0.6$. Recall that the standard definition of a dark matter halo is an object with a mean interior density of 200 times the background density. Second, the gaseous extent scaled with the halo mass as $M_h^{t}$ and $t=1/3$, following the expectation of the scaling relation between halo mass and the corresponding virial radius. The results presented in \\S\\ 4 offer a direct test of these assumptions. To compare the empirical constraints of Equations (6) \\& (7) with our initial model assumptions, we first derive the corresponding dark matter halo mass $M_h$ and their halo radius $R_{200}$ for galaxies of known absolute $B$-band magnitudes. The adopted $M_{\\rm B_*}$ for the analysis in \\S\\ 4 corresponds to a magnitude of $M_{\\rm b_J}-5\\,\\log\\,h=-19.7$ in the bandpass of the Two-Degree Field Galaxy Redshift Survey (2dFGRS, Colless \\etal\\ 2001). The 2dFGRS probes galaxies at $z\\sim 0.1$. Excluding galaxies that exist as satellites in a cluster environment\\footnote{We note that the fraction of satellite galaxies is $\\lesssim 15$ \\% for $L_*$ galaxies (Tinker \\etal\\ 2007; ZCZ07). In addition, none of the galaxies in our sample are observed to be part of a larger group or cluster. Here we carry out our calculations, assuming that most of galaxies in our sample are central galaxies in their dark matter halos.}, the mean halo mass for $M_{\\rm B_*}$ galaxies is $10^{12.5}\\,h^{-1}\\,\\hmsol$ at $z\\sim 0.1$ (Tinker \\etal\\ 2007; van den Bosch \\etal\\ 2007). At $z\\sim 1$, galaxies of this magnitude on average reside in $10^{11.9}\\,\\hmsol$ halos (ZCZ07). Interpolating in $\\log\\,(1+z)$, we find that galaxies of $M_{\\rm B_*}-5\\,\\log\\,h=-19.8$ are expected to reside in halos of $M_h=10^{12.3}\\,\\hmsol$ at $z=0.4$ and that the halo radius for $M_{\\rm B_*}$ galaxies is $R_{200}=212\\,h^{-1}$ physical kpc. The best-fit $R_{\\rm gas*}$ from Equation (7) therefore implies $R_{\\rm gas}\\approx 0.4\\,R_{200}$. Next, we derive the expected scaling relation between $R_{\\rm gas}$ and halo mass $M_h$. The halo occupation analysis of 2dFGRS galaxies from Tinker \\etal\\ (2007) showed that the relationship between $M_h$ and luminosity for galaxies of $L_{b_J}\\lesssim L_{*}$ at $z\\sim 0.1$ is $M_h = 10^{12.5}\\,(L_B/L_{B_*})^{1.3}\\,\\hmsol$. this monotonic relationship is appropriate for galaxies that reside at the {\\it centers} of their dark matter halos and are the brightest galaxy in the halo. A similar scaling relation was obtained from DEEP2 data by ZCZ07 for galaxies at $z\\sim 1$. Because there is a monotonic relationship between halo mass and galaxy luminosity, the scaling relation of $R_{\\rm gas}$ with $L_B$ in Equation (2) is equivalent to a scaling relation between $M_h$ and $R_{200}$ that are related by $M_h\\propto R_{200}^3$. Adopting $\\beta=0.35$ from Equation (6), we derive $R_{\\rm gas}=R_{\\rm gas*}\\times [M_h/(10^{12.3}\\,\\hmsol)]^{0.3}$ $h^{-1}$ kpc at $z\\sim 0.4$. The calculations above show that observations support the initial assumptions in TC08 that the gaseous radius is a constant fraction of the halo radius for all halos and that the gaseous radius scales with halo mass according to $M_h^{t}$ with $t\\approx 1/3$. We note that our galaxy sample therefore probes halo mass range from $10^{10.5}\\,\\hmsol$ to $10^{12.8}\\,\\hmsol$, and that the best-fit scaling relation is driven by galaxies in the cold-hot transition regime. Next, we compare the observed covering fraction of Mg\\,II absorbers with the predicted mean of TC08. As summarized in \\S\\ 1, we find in the halo occupation analysis of TC08 that dark matter halos of $M_h=10^{11.5-12.5}$ \\hmsol\\ on average have a covering fraction of unity $\\langle\\kappa(M_h)\\rangle=1$ at $R_{\\rm gas}\\le R_{\\rm 200}/3$ for Mg$^+$ and $\\langle\\kappa(M_h)\\rangle=0$ at larger radii. This halo mass range corresponds to a luminosity range of $L_B=0.2-1.4\\,L_{B_*}$ at $z\\sim 0.4$, following the mass-to-light scaling relation discussed above. Sixteen galaxies in our sample have $L_B\\ge 0.2\\,L_{B_*}$, eleven of which are at $\\log\\,\\rho +0.14\\times (M_B-M_{B_*}) < \\log\\,R_{\\rm gas*}$. We identify a corresponding Mg\\,II absorber for every one of these galaxies, indicating a mean covering fraction of $\\langle\\kappa\\rangle=100$ \\% at $R0.76$ with greater than 95\\% confidence level. Three of the seven remaining $L_B<0.2\\,L_{B_*}$ galaxies in our sample have $\\log\\,\\rho +0.14\\times (M_B-M_{B_*}) < \\log\\,R_{\\rm gas*}$. Two of them do not have corresponding Mg\\,II absorbers identified to a sensitive upper limit, implying $\\langle\\kappa\\rangle=33$ \\% at $R>k_\\bot$ interacting with other fast waves generate high-frequency Alfven-waves with $k_\\|>>k_\\bot$. We expect that the scattering by thus generated Alfven modes to be similar to scattering by fast modes that created them. Therefore, within the simplified approach adopted in the paper, we do not consider this type of interactions.} where the velocity scales as $v_k\\propto k^{-1/4}$ and in each wave-wave collision a small fraction of energy equal to $v_{ph}/v_k$ is transferred to smaller scales. In low $\\beta$ medium, the three dimensional energy spectrum is (YL02) \\bea W(k)= \\frac{nm_i\\delta V^2}{8\\pi}k^{-\\frac{7}{2}}L^{-\\frac{1}{2}}\\left(\\begin{array}{cc}k_ik_j/k_\\bot^2&\\\\ &0\\end{array}\\right), \\label{fastspec} \\eea with a cascade time scale of \\be \\tau_{cas}=(v_{ph}/v_k)(kv_{k})^{-1}=(L/\\delta V)M_A^{-1}(kL)^{-1/2}. \\label{tcasfast} \\ee Here $k_{i,j}$ refers to the x,y components of the wave vector ${\\bf k}$, $n$ is the density of the plasma with ions of mass $m_i$ ($\\sim$ proton mass), $\\delta V$ is the initial perturbation injected at the outer scale $L\\equiv k_{\\rm min}$), $v_{ph}=\\omega/k$ is the phase speed of fast mode with frequency $\\omega$ and wave vector $k$, and $M_A=\\delta V/v_A$ is the Alfv\\'enic Mach number for the Alfv\\'en speed $\\simeq v_A=B/\\sqrt{4\\pi nm_i}$. On small scales, the spectrum of turbulence is affected by damping. Damping becomes important at a wave vector $k_c$ when the damping time $\\Gamma_{k}^{-1}$ becomes comparable or shorter than the cascading time $\\tau_{cas}$. Beyond this wavevector the turbulence spectrum falls off rapidly. In fully ionized plasma, there are basically two kinds of damping. Collisional damping is important on scales greater than the Coulomb collision mean free path $\\lambda_{\\rm Coul}$. In solar flares $\\lambda_{\\rm Coul}\\sim 5\\times 10^{7}{\\rm cm}\\left(\\frac{T}{10^7{\\rm K}}\\right)^{2}\\left(\\frac{10^{10}{\\rm cm}^{-3}}{n}\\right)$ and the relevant scales are shorter so that collisionless damping is dominant. Taking into account interactions with thermal and nonthermal particles, Petrosian, Yan \\& Lazarian (2006, henceforth PYL06) studied damping of fast modes in solar corona condition and showed that most the damping is also highly anisotropic. For most angles of propagation the inertial range is truncated at large scales. Only quasi-parallel and quasi-perpendicular waves reach short scales comparable to ion gyroradius% \\footnote{The quasi-perpendicular most likely will be damped because of magnetic field wanderings (see PYL06).}. For plasma $\\beta_p=P_{gas}/P_{mag}\\lesssim 0.1$, the damping due to electrons dominates as it's easier for electrons to catch up with the waves. The corresponding damping rate is (Ginzburg 1961, YL02) \\begin{eqnarray} \\Gamma_{L} & = & \\frac{\\sqrt{\\pi\\alpha\\beta_p}}{2}\\omega\\frac{\\sin^{2}\\theta}{\\cos\\theta}\\exp\\left(-\\frac{\\alpha}{\\beta_p\\cos^2\\theta}\\right), \\label{Ginz} \\end{eqnarray} where $\\alpha=m_e/m_H$, $\\theta$ is the angle between the wave vector and the magnetic field. By equating the above equation and eq.(\\ref{tcasfast}), we obtain the cutoff scale of turbulence owing to the collisionless damping, \\be k_c L=\\frac{4M_A^4\\cos^2\\theta}{\\pi\\alpha\\beta_p\\sin^4\\theta}\\exp\\left(\\frac{2\\alpha} {\\beta_p\\cos^2\\theta}\\right). \\label{landauk} \\ee In what follows we shall use this spectrum of fast modes to calculate the acceleration and confinement of particles by fast modes in solar flare conditions. \\begin{table} \\caption{Notations in this paper.} \\begin{tabular}{|l l|} \\hline a & power law index of nonthermal particle distribution\\\\ A(E) & acceleration rate\\\\ $E_0$ & lower energy limit of nonthermal particles\\\\ {\\bf k}& wave vector\\\\ $\\omega$ & wave frequency\\\\ $k_c$& turbulence cutoff wavenumber due to damping\\\\ L& energy injection scale\\\\ $\\delta V$& injection speed\\\\ $M_A$ & Alfv\\'enic Mach number $\\delta V/v_A$\\\\ n & number density of the corona\\\\ N(E) & number density of nonthermal particles per energy bin\\\\ $N_0$ & N(E) at the lower energy limit $E_0$\\\\ T & corona temperature\\\\ $v_A$ & Alfv\\'en speed\\\\ v & particle speed\\\\ $\\mu$ & cosine of particle pitch angle\\\\ W({\\bf k})& turbulence spectral energy density\\\\ $\\alpha$& ratio of electron mass $m_e$ to ion mass $m_i$\\\\ B& magnetic field\\\\ $\\beta_p$& ratio of gas pressure to magnetic pressure\\\\ $\\beta$& v/c\\\\ $\\beta_A$& $v_A/c$\\\\ $\\Gamma$ & wave damping rate\\\\ $\\tau_{cas}$ & cascading time of the turbulence\\\\ $\\tau_{acc}$ & acceleration time scale\\\\ $\\tau_{loss}$ & energy loss time scale of particles\\\\ $\\tau_{esp}$ & escaping time of CRs from the system\\\\ $\\eta$ &$\\cos\\theta$\\\\ $\\eta_c$ &$v_A/v$\\\\ $\\lambda_{Coul}$& Coulomb collisional mean free path\\\\ $\\lambda_\\|$ ¶llel mean free path of CRs\\\\ $\\theta$& pitch angle of a fast modes\\\\ $\\delta \\theta$& variation of $\\theta$ in turbulence\\\\ \\hline \\end{tabular} \\end{table} \\begin{table} \\caption{The physical parameters of solar flares we adopted.} \\begin{tabular}{|c|c|c|c|c|} \\hline T(KeV)&n(cm$^{-3}$)&$\\beta_p$&L(cm)&$M_A$ \\\\ \\hline 1&$10^{10}$&$0.01, 0.04, 0.1$&$10^9$&$0.3$\\\\ \\hline \\end{tabular} \\end{table} In view of various difficulties with quasilinear theory (QLT), a number of nonlinear theories (NLT) have been proposed (see Dupree 1966; V\\\"olk 1973; Jones, Kaiser \\& Birmingham 1973; Goldstein 1976; Felice \\& Kulsrud 2001; Matthaeus et al. 2003; Shalchi 2005). Based on particle trapping due to large scale magnetic perturbations (V\\\"olk 1975), we developed a nonlinear formalism in YL08 to treat cosmic ray scattering in MHD turbulence. In view of these progresses, we believe the time is ripe to investigate the stochastic acceleration by the tested model of MHD turbulence in solar flares. In \\S 2 we describe how the nonlinear effects are formulated. In \\S 3 and \\S 4 we present results on acceleration and confinement of the particles, and in \\S 5 and \\S 6 we present a brief discussion and summary of the results. Some mathematical details are given in the appendix. ", "conclusions": "In this paper we have discussed the effects that the cascade of turbulence has on acceleration and heating of Solar corona. There the energy is injected at large scales much larger than any plasma scale concerned and this justifies a magnetohydrodynamic treatment of the large scale motions. Due to recent insights into the physics of MHD cascade and its interaction with charged particles we reduced a complex problem of acceleration and heating to a more manageable problems of interactions of Alfv\\'en, slow and fast modes with plasma and energetic particles. Alfv\\'enic turbulence is inefficient in scattering and accelerating particles because of its anisotropy (Chandran 2000, YL02). Fast modes, instead, have been identified as the MHD turbulence modes dominating the interaction with cosmic rays (YL02,YL04). In this paper, we apply the result to the stochastic acceleration in solar flares. We assume that the MHD turbulence is strong, i.e. that the critical balanced condition is satisfied for Alfv\\'enic modes. Therefore, to describe fast mode of MHD turbulence we appeal to the results by Cho \\& Lazarian (2002, 2003) on the scaling and coupling of Alfv\\'enic and fast modes. The corresponding papers claim the isotropy of the fast modes. One may expect to see deviations from isotropy, however. For instance, slab Alfv\\'en modes created by streaming instabilities are subject to non-linear damping by the ambient Alfv\\'enic turbulence (YL02, YL04, Farmer \\& Goldreich 2004, Lazarian \\& Beresnyak 2006). Beresnyak \\& Lazarian (2008) showed that the corresponding damping of the quasi-parallel modes depends on the angle between their ${\\bf k}$ vector and the direction of magnetic field. One might expect quasi-slab fast mode to be subject to a similar damping by strong Alfv\\'enic turbulence. This, however, has not been demonstrated so far. If at the injection scale $\\delta B\\ll B$, the Alfv\\'enic turbulence is weak (see Galtier et al 2000) and develop a cascade with $k_{\\|}=const$. The interaction of the weak Alfv\\'enic turbulence with other modes can be very different from that of the strong Alfv\\'enic turbulence. For instance, it was shown by Chandran (2005) that fast modes develop anisotropy (i.e. the energy in the quasi-slab modes is reduced) owing to their interaction with Alfv\\'en modes in the weak regime. However, such the Alfvenic weak turbulence has a limited inertial range (see discussion in Cho, Lazarian \\& Vishniac 2003) and at sufficiently large $k$ transfers into a strong turbulence. Moreover, magnetic reconnection in Solar Flares should produce perturbations $\\delta B\\sim B$, which should induce strong turbulence from the very beginning. Strong Alfv\\'enic turbulence is characterized by $k_{\\bot}\\gg k_{\\|}$ with the GS95 relation $k_{\\|}\\sim k_{\\bot}$ defining a cone in the Fourier space where most of the turbulent energy resides. However, the aforementioned relation should not be understood too literally. The energy outside the cone is not zero and the modes with $k_{\\|}>k_{\\bot}$ are present. Such Alfv\\'enic modes are weakly interacting even being a part of the strong Alfv\\'enic turbulence. Therefore, the Chandran (2005) model is applicable to them. Nevertheless, according to Cho, Lazarian \\& Vishniac (2002) the energy in these modes is exponentially reduced\\footnote{Because of this exponential reduction the rates of Alfv\\'enic scattering in YL02, which was the first study to take into account the modes outside the GS95 cone, were still grossly subdominant to the fast modes.}. Therefore, assuming that the Alfv\\'enic turbulence is injected at a scale much larger than the typical gyroradius of the energetic particles, we did not consider the anisotropies, introduced by the process of the fast wave cascading induced by Alfvenic modes with large $k_{\\|}$ and small $k_{\\bot}$ (cf. Chandran 2005).\" The model for MHD turbulence that we adopted in the paper is the turbulence where the back-reaction of energetic particles is limited to changing the cut-off of the turbulence. A more fundamental modifications of turbulence are conceivable, however. For instance, Lazarian \\& Beresnyak (2006) argued that compressions of the fluid of energetic particles may result in the gyroresonance instabilities that can induce an additional quasi-parallel component of the Alfvenic waves. If true, these waves would interact with fast modes in the manner described by Chandran (2005), which would affect the fast mode isotropy. However, if substantial portion of energy resides within these quasi-slab modes, their major effect will be the direct gyroresonance acceleration of energetic particles. We felt that the modification of MHD turbulence arising from the instabilities within energetic particles, e.g. by the process in Lazarian \\& Beresnyak (2006), is beyond the scope of the present paper. Apart from the issue of isotropy, there are potential issues related to the exact scaling of fast modes. One dimensional numerical simulations in Suzuki, Lazarian \\& Beresnyak (2007) indicate that fast modes may develop a shock-like cascade, which differs from the finding in 3D MHD calculations in Cho \\& Lazarian (2002, 2003). We adopted the model from the latter works, but wait for higher resolution 3D numerical runs to rectify the scaling. In addition to being strong, MHD turbulence that we considered was balanced in the sense that the equal flux of energy was assumed in every possible direction. The properties of imbalanced turbulence (see Cho, Lazarian \\& Vishniac 2002; Lithwick, Goldreich \\& Sridhar 2007; Beresnyak \\& Lazarian 2007, Chandran 2008) can be very different from the balanced one. However, we expect the flow of Alfv\\'en waves, which constitute the weak turbulence, to be subjected to reflection within a Solar corona environment that we deal with. As the result we expect only marginal imbalance for the problem that we deal with. A threshold for the TTD acceleration is $v\\gtrsim v_A$ set by the Cherenkov resonance condition. In the low $\\beta_p$ environment, the thermal protons can not be accelerated. The low energy threshold would be $\\sim 2(T/10^7)/\\beta_p$KeV. Thermal electrons, instead, can have TTD interactions unless the plasma beta is too low $\\beta_p\\lesssim m_e/m_p$. Our calculations show that TTD acceleration dominates over gyroresonance for the energy range we consider. The acceleration efficiency decreases with the plasma $\\beta_p$ as damping of the fast modes increase with $\\beta_p$. Particle acceleration rate depends on the wave spectrum and the wave damping rate are partially determined by the particle spectrum. In general, it is required that a self-consistent treatment of the evolution of turbulence and particle acceleration. The calculations above assumes that the spectrum of the turbulence is determined by the cascade and damping by thermal particles. Our numerical calculation of these integrals shows that up to $\\sim$ 10\\% of total energy of turbulence is being transferred to nonthermal particles for the given set of parameters. In most cases, the back reaction of nonthermal particles is negligible as the damping due to the interaction with these nonthermal particles is smaller than thermal damping taking into account field line wandering. In some large flares, however, a large amount of particles need to be accelerated from the thermal reservoir to energies $\\gg k_BT$ (PYL06). In this case, the damping by nonthermal particles can be significant and one needs to solve the coupled equation of the evolution of turbulence and particles simultaneously. Isotropization is important for the process of acceleration. If there is not enough isotropization, the acceleration by TTD will stop quickly as only parallel velocity is increased during the process. Moreover, without enough scattering, particles will leave the system before they get accelerated. Scattering by fast modes is shown to be adequate for isotropization of protons. Different from the acceleration, the scattering by gyroresonance, even smaller than the TTD scattering rate, plays an essential role in determining the scattering of parallel moving particles, which can be a substantial portion of the particles because of the TTD acceleration. For electrons, due to their small gyro-radii, gyroresonance does not occur with MHD turbulence except for those high energy electrons. The actual scattering can happen with the plasma turbulence. The work by Cho \\& Lazarian (2004) shows that whistler modes are even more anisotropic than Alfv\\'en modes. We know gyroresonance is very inefficient with anisotropic turbulence. Analogous to the MHD regime, parallel propagating modes may be generated by kinetic instabilities (see Tsytovich 1977) and be a candidate for interaction with electrons. The TTD interaction itself may induce gyroresonance instability through creating an anisotropic distribution of particles with respect to the magnetic field (Lazarian \\& Beresnyak 2006). Our estimate shows that the whistler waves (for moderate energy) and Alfv\\'en wave (for high energy) generated by the anisotropy instability can provide effective scattering and isotropization for the electron acceleration. Acceleration of particles by fast modes were previously studied by a number of authors, including Miller, Larosa \\& Moore (1996), Schlickeiser \\& Miller (1998). In their studies, turbulence is assumed isotropic with either Kolmogorov or Kraichnan spectrum. Although coupled equations of wave and particle evolutions were solved in Miller, Larosa \\& Moore (1996), we feel that the one dimensional treatment they adopt is problematic, as the damping caused by the TTD interaction with particles is anisotropic. In this paper, we start with a more physically motivated and numerically tested of turbulence. We deal with fast modes, which are subject to much more efficient linear dissipation. On the small scales, however, because of the dissipation above, fast modes develop anisotropy. In our treatment, this anisotropy is strongly affected by field line wandering which is determined by the Alfv\\'enic Mach number and plasma $\\beta$ and the efficiency of the acceleration is substantially influenced accordingly. In addition, scattering and confinement are treated using the nonlinear theory we have recently developed (Yan \\& Lazarian 2008). We believe that the approach that we developed is applicable beyond pure Solar Flare problems, e.g. it can be modified to study turbulent acceleration in the medium within clusters of galaxies (see Brunetti \\& Lazarian 2007)." }, "0801/0801.4029_arXiv.txt": { "abstract": "We present the results from a multiwavelength campaign on the TeV blazar 1ES 1959+650, performed in May, 2006. Data from the optical, UV, soft- and hard-X-ray and very high energy (VHE) gamma-ray (${\\rm E} > 100$ GeV) bands were obtained with the \\suzaku and \\swift satellites, with the \\magic telescope and other ground based facilities. The source spectral energy distribution (SED), derived from \\suzaku and \\magic observations at the end of May 2006, shows the usual double hump shape, with the synchrotron peak at a higher flux level than the Compton peak. With respect to historical values, during our campaign the source exhibited a relatively high state in X-rays and optical, while in the VHE band it was at one of the lowest level so far recorded. We also monitored the source for flux-spectral variability on a time window of 10 days in the optical-UV and X-ray bands and 7 days in the VHE band. The source varies more in the X-ray, than in the optical band, with the 2-10 keV X-ray flux varying by a factor of $\\sim 2$. The synchrotron peak is located in the X-ray band and moves to higher energies as the source gets brighter, with the X-ray fluxes above it varying more rapidly than the X-ray fluxes at lower energies. The variability behaviour observed in the X-ray band cannot be produced by emitting regions varying independently, and suggests instead some sort of ``standing shock'' scenario. The overall SED is well represented by an homogeneous one-zone synchrotron inverse Compton emission model, from which we derive physical parameters that are typical of high energy peaked blazars. ", "introduction": "It is widely accepted that the spectral energy distribution (SED) of blazars is dominated by a non-thermal continuum, produced within a relativistic jet closely aligned with the line of sight, making these objects very good laboratories to study the physics of relativistic jets. The overall emission, from radio to $\\gamma$--rays and, in some cases, to the multi--TeV band, shows the presence of two well--defined broad components (von Montigny et al. 1995; Fossati et al. 1998). Usually, for the blazars that are detected in the TeV bands, the first components peaks in the UV -- soft-X--ray bands (HBL: high energy peaked blazars, Padovani \\& Giommi (1995) and the second one in the GeV--TeV region. The blazar emission is very successfully interpreted so far in the framework of Synchrotron Inverse Compton models. The lower energy peak is unanimously attributed to synchrotron emission by relativistic electrons in the jet, while the second component is commonly believed to be Inverse Compton emission (IC) from the same electron population (e.g., Ghisellini et al. 1998), although different scenarios have been proposed (e.g. B\\\"ottcher 2007). Since the discovery of the first blazar emitting TeV radiation, Mrk~421 (Punch et al. 1992), TeV blazars have been the target of intense observational and theoretical investigations. Indeed, the possibility of coupling observations of the emission produced by very high energy electrons, in the VHE (very high energy) band (up to Lorentz factors of the order of $10^7$), with observations in the soft and hard X--ray bands offers a unique tool to probe the cooling and acceleration processes of relativistic particles. In fact the synchrotron peak of these sources is usually located in the soft X--ray, while it is in the hard X--ray band that the synchrotron emission by the most energetic electrons can be studied and that the low energy part of the Compton emission component can start to dominate. Studies conducted simultaneously in the soft and hard X--ray and in the VHE bands are of particular importance, since in the simple Synchrotron-Self Compton (SSC) framework one expects that variations in X--rays and TeV should be closely correlated, being produced by the same electrons (e.g. Tavecchio et al. 1998). In fact even the first observations at X--ray and TeV energies yielded significant evidence of correlated and simultaneous variability of the TeV and X--ray fluxes (Buckley et al. 1996; Catanese et al. 1997). During the X-ray/TeV 1998 campaign on Mrk\\,421 a rapid flare was detected both at X--ray and TeV energies (Maraschi et al. 1999). Subsequent observations confirmed these first evidences also in other sources. Note however that the correlation seems to be violated in some cases, as indicated by the observation of an ``orphan'' (i.e. not accompanied by a corresponding X--ray flare) TeV event in 1ES 1959+650 (Krawczynski et al. 2004). In the case of PKS\\,2155-304 a giant TeV flare recorded by HESS (Aharonian et al. 2007), with a TeV flux \"night-average\" intensity of a factor of $\\sim 17$ larger than those of previous campaigns, was accompanied by an increase of the X-ray flux of only a factor of five without a significant change of the X-ray spectrum (Foschini et al. 2007). In the one-zone SSC scenario this can be accomplished with an increase of the Doppler factor and the associated relativistic electrons together with a decrease of the magnetic field. Therefore, it is important to obtain simultaneous observations over the largest possible UV and X-ray range together with simultaneous VHE observation to probe the correlation between the synchrotron and VHE emission. To this end we organised a multiwavelength campaign to observe the blazar 1ES\\,1959+650 in the optical, UV, soft and hard X-ray up to the VHE gamma-ray (${\\rm E} > 100$ GeV) bands. This is a bright and flaring X-ray and VHE source that has already been observed many times in these bands. It was discovered in the radio band as part of a 4.85 GHz survey performed with the 91 m NRAO Green Bank telescope (Gregory \\& Condon 1991; Becker, White \\& Edwards 1991). In the optical band it is highly variable and shows a complex structure composed by an elliptical galaxy (M$_R =-23$, $z$=0.048) plus a disc and an absorption dust lane (Heidt et al. 1999). The mass of the central black hole has been estimated to be in the range $1.3-4.4 \\times 10^8 \\, {\\rm M_\\odot}$ as derived either from the stellar velocity dispersion or from the bulge luminosity (Falomo et al. 2002). The first X-ray measurement was performed by {\\it Einstein}-IPC during the Slew Survey (Elvis et al. 1992). Subsequently, the source was observed by \\rosat, {\\it Beppo}SAX, {\\it RXTE, ARGOS, XMM-Newton}. In particular two {\\it Beppo}SAX pointings, simultaneous with optical observations, were triggered in May-June 2002 because the source was in a high X-ray state. These data showed that the synchrotron peak was in the range 0.1-0.7 keV and that the overall optical and X-ray spectrum up to 45 keV was due to synchrotron emission with the peak moving to higher energy with the higher flux (Tagliaferri et al. 2003). The overall SED, with non simultaneous VHE data could be modelled with a homogeneous, one-zone synchrotron inverse Compton model. The results of a multiwavelength campaign performed in May-June 2003 are presented in Gutierrez et al. (2006). This campaign was triggered by the active state of the source in the X-ray band and it was found that the X-ray flux and X-ray photon index are correlated. A similar result was found by Giebels et al. (2002) using{\\it RXTE} and {\\it ARGOS} data. This correlation shows that the X-ray spectrum in the 1-16 keV band is harder when the source is brighter. In the VHE band the source was detected by the HEGRA, Whipple and MAGIC telescopes (Aharonian et al. 2003; Holder et al. 2003a, Albert et al. 2006). One of the most important results of these observations is probably the ``orphan\" flare mentioned above, seen in the VHE band and not in X-rays (Krawczynski et al. 2004). 1ES\\,1959+650 is therefore one of the most interesting and frequently observed high energy sources of recent years. With the aim of obtaining a better description of the broad band X-ray continuum and in particular of observing simultaneously the synchrotron and IC components, we asked for simultaneous \\suzaku and \\magic observations that were carried out in May 23-25, 2006. Around the same epoch we obtained various Target of Opportunity (ToO) short pointings with {\\it SWIFT} and observed the source also in the optical R-band from ground. A preliminary analysis of these data is reported in Hayashida et al. (2007). In the following we report the data analysis (Sec.2) and the results (Sec.3). The discussion and conclusions are given in Sec. 4, where we model the SED in the framework of a homogeneous, one-zone SSC model. Throughout this work we use $H_{\\rm 0}\\rm =70\\; km\\; s^{-1}\\; Mpc^{-1} $, $\\Omega _{\\Lambda}=0.7$, $\\Omega_{\\rm M} = 0.3$. ", "conclusions": "The full SED of 1ES 1959+650 as measured at the end of May 2006 is reported in Fig.~\\ref{sed}, together with other historical data. During our multiwavelength campaign we simultaneously observed the SED from the optical, to the UV, soft and hard X-rays and VHE bands, monitoring both the synchrotron and Compton components. The historical data in this figure show very strong changes in the X-ray band, while in the optical this is much more attenuated. A behaviour that is also found in the results obtained from our observing campaign. During our multiwavelegth campaign the source is found to be in a high state with respect to the historical behaviour both in X-ray and optical (e.g. Tagliaferri et al. 2003 and Fig. \\ref{lc_optical}), although not at the highest state as observed in the X-ray (e.g. Holder et al. 2003b, see Fig. \\ref{sed}). In the VHE band, instead, the source is at one of the lowest state so far recorded. We also found that the X-ray fluxes at energies above the synchrotron peak vary more rapidly than the X-ray fluxes below the peak. Also the VHE band shows historical strong variability, in particular if we consider that in this band there are fewer observations than in the optical or X-ray ones. However, from our data we do not see strong (i.e. a factor of 2-3) variability in the VHE band. Our \\magic data are probably monitoring the part of the SED slightly above the peak of the Compton component. Therefore, one would expect to see a high level of variability. The lack of variability in the \\magic data and the low flux level recorded both indicate that the source was not very active in this band. Overall we can say that during our campaign the source was quite stable (i.e. did not vary by more than a factor of 2) from the optical to the VHE band. The observed X--ray variability behaviour allows a few interesting considerations about the properties of the emitting regions. First, note that the variability is not random, but follows a raising/decay trend on a timescale of $\\sim$10 days (see the \\swift-XRT results). In this observed time $\\Delta t$, a single blob moving with a bulk Lorentz factor $\\Gamma\\sim 18$ (see below) moves by a distance $\\Delta z \\sim c\\Delta t \\Gamma^2\\sim 2.7$ pc. Therefore we cannot assume that we are observing a single moving blob travelling that far, since the blob would expand, loose energy by adiabatic losses, and change (decrease) its magnetic field. This in turn would decrease the produced flux and would lengthen the variability timescale. Also the internal shock model (Spada et al. 2001, Guetta et al. 2004) can not explain the variability we are observing. In fact, in this model the radiation is produced in a shock resulting from the collision of two shells moving at slightly different velocities. In this case the variability is predicted to be erratic, therefore to explain the variability we are seen we have to finely tune the different $\\Gamma$ of the shells. We are thus led to consider the possibility that the observed radiation originates in the same region of the jet, through some kind of ``standing shock\". For instance, we might think to the interaction of a fast spine and a shear layer occurring at about the same distance from the central powerhouse (see Ghisellini, Tavecchio \\& Chiaberge 2005 for mode details, including the possibility of radiative deceleration of the spine through the ``Compton rocket\" effect in TeV blazars). A ``standing shock\" scenario has already been proposed by Krawczynski et al. (2002) in order to explain the tight correlation between X-ray and TeV flares observed in Mrk501 and it is discussed in some detail also by Sokolov et al.( 2004). As we did with the previous multiwavelength observing campaigns on 1ES\\,1959+650 that we organised based on the {\\it Beppo}SAX observations (Tagliaferri et al. 2003), we can try to fit our SED with a homogeneous, one-zone synchrotron inverse Compton model. During the {\\it Beppo}SAX campaigns, in order to derive the SSC physical parameters we had to assume a value for the Compton component, that we derived by rescaling a non-simultaneous VHE spectrum based on the X-ray flux. This time we have also the VHE observations, therefore both SSC components are constrained by real data. As shown by Fig. \\ref{sed}, the X-ray spectrum as observed by \\suzaku and \\swift is about a factor of 2 higher than the one measured with \\sax and also the synchrotron peak has moved to somewhat higher energy, confirming the previous results of a higher energy peak with higher fluxes (e.g. Tagliaferri et al. 2003), that is typical for HBLs (see the dramatic case of MKN\\,501, Pian et al. 1998). The optical fluxes are similar to the one reported for the 2002 SED. The observed VHE spectrum is similar, but lower, to that one of the 2002 SED. In summary the 2006 SED has optical fluxes that are similar to that ones of the 2002, the X-ray fluxes are a factor of 2 higher and the VHE fluxes are a factor of $\\sim 2$ lower. In the assumed one-zone SSC model, the source is a sphere with radius $R$ moving with bulk Lorentz factor $\\Gamma$ and seen at an angle $\\theta$ by the observer, resulting in a Doppler factor $\\delta$. The magnetic field is tangled and uniform while the injected relativistic particle are assumed to have a (smooth) broken power law spectrum with normalisation $K$, extending from $\\gamma _{\\rm min}$ to $\\gamma _{\\rm max}$ and with indexes $n_1$ and $n_2$ below and above the break at $\\gamma _b$. Assuming this model, the SED of May, 2006 can be well represented using the following parameters: $\\delta=18$, R$=7.3 \\times 10^{15}$ cm, B=0.25 G, K=$2.2 \\times 10^3$ cm$^{-3}$ and an electron distribution extending from $\\gamma _{\\rm min} = 1$ to $\\gamma _{\\rm max} = 6.0 \\times 10^5$, with a break at $\\gamma _{\\rm b}= 5.7 \\times 10^4$ and slopes $n_1=2$ and $n_2=3.4$. The intrinsic luminosity is $L' = 5.5 \\times 10^{40}$ erg s$^{-1}$. If we compare these values with the one we derived for the 2002 SED (though in that case we use a slightly different emission model), we saw that the parameters are very similar, with a source that is slightly more compact, a lower magnetic field and an almost identical Doppler factor. Similar values are also found to explain the SED of PKS2155-304 during and after the strong TeV flare observed in July, 2006; although in that case we found less steep slopes for the electrons and an higher value of $\\delta$ (see Foschini et al. 2007). Once again, the physical parameters that we derived assuming a one-zone SSC model are typical of HBL objects. Finally, the historical SEDs of 1ES1959+650 shows that in this source the synchrotron emission is dominating above the Compton one." }, "0801/0801.1110_arXiv.txt": { "abstract": "Using the Gemini Near-InfraRed Spectrograph (GNIRS), we have completed a near-infrared spectroscopic survey for $K$-bright galaxies at $z\\sim2.3$, selected from the MUSYC survey. We derived spectroscopic redshifts from emission lines or from continuum features and shapes for all 36 observed galaxies. The continuum redshifts are driven by the Balmer/4000\\,\\AA\\ break, and have an uncertainty in $\\Delta{}z/(1+z)$ of $<0.019$. We use this unique sample to determine, for the first time, how accurately redshifts and other properties of massive high-redshift galaxies can be determined from broadband photometric data alone. We find that the photometric redshifts of the galaxies in our sample have a systematic error of 0.08 and a random error of 0.13 in $\\Delta{}z/(1+z)$. The systematic error can be reduced by using optimal templates and deep photometry; the random error, however, will be hard to reduce below 5\\%. The spectra lead to significantly improved constraints for stellar population parameters. For most quantities this improvement is about equally driven by the higher spectral resolution and by the much reduced redshift uncertainty. Properties such as the age, $A_V$, current star formation rate, and the star formation history are generally very poorly constrained with broadband data alone. Interestingly stellar masses and mass-to-light ratios are among the most stable parameters from broadband data. Nevertheless, photometric studies may overestimate the number of massive galaxies at $210^{11}\\,M_{\\odot}$ at $z\\sim2.5$) will not be feasible in the foreseeable future. Until the next generation of space missions and $>20$\\,m ground-based telescopes we remain largely dependent on photometric redshifts for studies of large and faint galaxy samples beyond $z>1.5$. Our provisional dependency on broadband photometric studies requires a more accurate calibration and understanding of the involved systematics. The current spectroscopic samples used for calibration of photometric high-redshift studies are based primarily on optical spectroscopy. As these samples are biased towards un-obscured star-forming galaxies, their calibration may not be representative for the total sample of massive galaxies. Photometric properties of red, massive galaxies at high redshift are poorly calibrated, and since red galaxies dominate the high mass end at $210^{11} M_{\\odot}$. The distribution of observed $J-K$, $R-K$ and photometric rest-frame $U-V$ colors are similar as those of a photometric mass-limited sample extracted from the deep MUSYC fields. This suggests that our spectroscopic sample is representative for a mass-limited ($>10^{11} M_{\\odot}$) sample at $210^{11}M_{\\odot}$) at $2.02$, larger spectroscopic samples over a larger redshift range are needed to fully map the systematics and accurately calibrate photometric studies. Furthermore, this study only applies to the high-mass end of the high-redshift galaxy population. Less massive galaxies may have other systematics, although that remains to be explored." }, "0801/0801.0927_arXiv.txt": { "abstract": "We present deep $I$ and $z'$ imaging of the colour-selected cluster RzCS 052 and study the color-magnitude relation of this cluster, its scatter, the morphological distribution on the red sequence, the luminosity and stellar mass functions of red galaxies and the cluster blue fraction. We find that the stellar populations of early type galaxies in this cluster are uniformly old and that their luminosity function does not show any sign of evolution other than the passive evolution of their stellar populations. We rule out a significant contribution from mergers in the buildup of the red sequence of RzCS 052. The cluster has a large ($\\sim 30\\% $) blue fraction and and we infer that the evolution of the blue galaxies is faster than an exponentially declining star formation model and that these objects have probably experienced starburst episodes. Mergers are unlikely to be the driver of the observed colour evolution, because of the measured constancy of the mass function, as derived from near-infrared photometry of 32 clusters, including RzCS 052, presented in a related paper. Mechanisms with clustercentric radial dependent efficiencies are disfavored as well, because of the observed constant blue fraction with clustercentric distance. ", "introduction": "The existence of a tight, and apparently universal, color-magnitude relation for galaxies in nearby clusters (e.g. Bower, Lucey \\& Ellis 1992; Andreon 2003; Lopez-Cruz, Barkhouse \\& Yee 2004; McIntosh, Rix \\& Caldwell 2005; Eisenhardt et al. 2007 and references therein) implies that the majority of the stellar populations of early-type cluster galaxies were formed at $z \\gg 1$ over relatively short timescales. Studies of clusters at high redshift, then, should allow us to witness the earlier stages of galaxy evolution, leading to the establishment of the present day luminosity function, color-magnitude relation and morphological mixtures. A classical example of this kind of studies is the detection of a blueing trend among galaxies in clusters at $z > 0.3$ by Butcher \\& Oemler (1984). Until large samples of high redshift ($z \\ga 0.8$) clusters become available, detailed `case studies' of individual objects may provide useful clues to the evolution of galaxy populations at half the Hubble time and beyond. Several studies have analyzed a number of such objects in detail (Blakeslee et al. 2003, 2006; Homeier et al. 2005, 2006; Holden et al. 2006; Mei et al. 2006a,b), using both ground-based imaging and spectroscopy and high-resolution imaging with the Advanced Camera for Surveys (ACS) on the Hubble Space Telescope (HST). These observations reiterate that the cluster early-type populations appear to be composed of old stellar populations which were probably in place at high redshift. Most of the studied high redshift clusters are selected from the X-ray catalogs, which may pre-select objects that have already formed a deep potential well. An alternative strategy is to use clusters selected via the prominent red sequence of early type galaxies (Gladders \\& Yee 2000). Several $z\\geq 1$ clusters have already been identified using the galaxy colours or their spectral energy distributions (Andreon et al. 2005; Stanford et al. 2005). Here we focus on RzCS 052 (J022143-0321.7), a rich (Abell class 2 or 3) cluster selected via a modified red sequence method, with a measured redshift of $1.016$ and a velocity dispersion of $710\\pm150$ km s$^{-1}$ and a modest X-ray luminosity of $(0.68 \\pm 0.47) \\times 10^{44}$ ergs s$^{-1}$ in the [1-4] keV band. Details about this objects and its properties may be found in Andreon et al. (2007). Because RzCS 052 is less X-ray luminous than clusters at similar redshift, and therefore does not possess a massive X-ray atmosphere, this object allows us to carry out a study of galaxy evolution in a different cluster environment and isolate the effects of gas on galaxy properties (Moran et al. 2007). The layout of the paper is as follows. We present the data reduction and analysis in the next section. Section 3 discusses the red sequence galaxies. Section 4 deals with the blue galaxies in RzCS 052. Finally, we summarize our results in Section 5. We adopt the concordance cosmological parameters $\\Omega_M=0.3$, $\\Omega_{\\Lambda}=0.7$ and $H_0=70$ km s$^{-1}$ Mpc$^{-1}$. Magnitudes are quoted in their native photometric system (Vega for $RI$, SDSS for $z'$ and instrumental for Megacam data). Results of our stochastical computations are quoted in the form $x\\pm\\sigma$ where $x$ is the posterior mean and $\\sigma$ is the posterior standard deviation. \\begin{figure} \\hbox{% \\psfig{figure=gal6.ps,width=2truecm,clip=}% \\psfig{figure=gal11.ps,width=2truecm,clip=}% \\psfig{figure=gal17.ps,width=2truecm,clip=}% } \\hbox{% \\psfig{figure=gal1.ps,width=2truecm,clip=}% \\psfig{figure=gal4.ps,width=2truecm,clip=}% \\psfig{figure=gal8.ps,width=2truecm,clip=}% \\psfig{figure=gal9.ps,width=2truecm,clip=}% } \\hbox{% \\psfig{figure=gal10.ps,width=2truecm,clip=}% \\psfig{figure=gal12.ps,width=2truecm,clip=}% \\psfig{figure=gal13.ps,width=2truecm,clip=}% \\psfig{figure=gal14.ps,width=2truecm,clip=}% } \\hbox{% \\psfig{figure=gal15.ps,width=2truecm,clip=}% \\psfig{figure=gal18.ps,width=2truecm,clip=}% } \\hbox{% \\psfig{figure=gal7.ps,width=2truecm,clip=}% \\psfig{figure=gal16.ps,width=2truecm,clip=}% \\hskip 2 truecm \\psfig{figure=gal5.ps,width=2truecm,clip=}% } \\caption[h]{Gallery of HST images of spectroscopically confirmed galaxies. The top row shows early-type galaxies. The second to fourth rows show red ($I-z'>-0.12$ mag) late-type galaxies. Last row shows blue ($I-z'<-0.12$ mag) late-type galaxies (two left-most panels) or unclassified galaxy (righmost panel). Each panel is $5''$ wide.} \\end{figure} ", "conclusions": "We presented new results on morphology, mass assembly history, role of environment and colour bimodality of galaxies in the colour selected, modest x-ray emitter, cluster RzCS 052 at $z=1.02$, as derived from VLT, HST, and optical data, supplemented by a coarser analysis of a large sample of about 45 clusters, from $z=0$ to $z=1.22$ (16 in the context of the evolution of red galaxies shown in Fig. 4, and 32 used to constraint the fate of the blue population in Sec 4). We found that the colour distribution of RzCS 052 is bimodal. Analysis of the morphological mix, slope and intercept of the colour-magnitude relation, and of the mass function shows that RzCS 052 red galaxies differs only by age from their local counterparts and that mergers play a minor role in building massive (down to~$M^*+2$) red galaxies in studied clusters, from $z=1$ to today. The situation is remarkably different for blue galaxies. The blue fraction, once accounted for the younger age of stellar populations at high redshift and for the higher star formation rate there, is larger in RzCS 052 than in nearby similar clusters, highlighting perhaps for the first time that something, in addition to the flow of the time, is making galaxies bluer at high redshift. Mergers are unlikely to be the driver of the observed colour evolution between $z=1$ and $z=0$, because of the measured constancy of the mass function, as derived from Spitzer photometry of 32 clusters in Andreon (2006). Mechanisms requiring a substantial intracluster medium, such as ram pressure stripping, are ruled out as well as direct driver, because of the very modest x-ray emission in RzCS 052. Mechanisms with a substantial different efficiency at the center and at one virial radius are strongly disfavored, because of the observed constant blue fraction." }, "0801/0801.2113_arXiv.txt": { "abstract": "We study the evolution of dwarf ($L_H$ $<$ 10$^{9.6}$ L$_{H \\odot}$) star forming and quiescent galaxies in the Virgo cluster by comparing their UV to radio centimetric properties to the predictions of multizone chemo-spectrophotometric models of galaxy evolution especially tuned to take into account the perturbations induced by the interaction with the cluster intergalactic medium. Our models simulate one or multiple ram pressure stripping events and galaxy starvation. Models predict that all star forming dwarf galaxies entering the cluster for the first time loose most, if not all, of their atomic gas content, quenching on short time scales ($\\leq$ 150 Myr) their activity of star formation. These dwarf galaxies soon become red and quiescent, gas metal-rich objects with spectrophotometric and structural properties similar to those of dwarf ellipticals. Young, low luminosity, high surface brightness star forming galaxies such as late-type spirals and BCDs are probably the progenitors of relatively massive dwarf ellipticals, while it is likely that low surface brightness magellanic irregulars evolve into very low surface brightness quiescent objects hardly detectable in ground based imaging surveys. The small number of dwarf galaxies with physical properties intermediate between those of star forming and quiescent systems is consistent with a rapid ($<$ 1 Gyr) transitional phase between the two dwarf galaxies populations. These results, combined with statistical considerations, are consistent with the idea that most of the dwarf ellipticals dominating the faint end of the Virgo luminosity function were initially star forming systems, accreted by the cluster and stripped of their gas by one or subsequent ram pressure stripping events. ", "introduction": "Dwarf galaxies ($M_B>$-18) are the most common objects in the nearby universe (Ferguson \\& Binggeli 1994). Their importance resides in the fact that they represent in cold dark matter models the building blocks of hierarchical galaxy formation (e.g. White \\& Rees 1978, White \\& Frenk 1991). Their study is thus fundamental for constraining models of galaxy formation and evolution.\\\\ Observations of dwarf galaxies, necessarily limited to the nearby universe, revealed however a more complex origin than that predicted by models, making this class of galaxies even more interesting. Among dwarf galaxies, dwarf ellipticals (dE) and the less luminous dwarf spheroidals (dSph)\\footnote{Unless specified, in the following we indicate with dE all quiescent dwarf galaxies, including dSph} are more common than star forming Im and BCDs (Ferguson \\& Binggeli 1994). Firstly thought as the low-luminosity extension of bright ellipticals, several observational evidences indicate that they form an independent class of objects (Bender et al. 1992). Dwarf and giant ellipticals have Sersic light profiles of index $1/n$, with $n$ progressively increasing with luminosity (Graham \\& Guzman 2003; Gavazzi et al. 2005a). The colour magnitude relations measured using resolved stars in local group dwarfs spheroidals (Mateo 1998; Grebel 1999), the optical (Conselice et al. 2003a) and UV (Boselli et al. 2005a) integrated color magnitude relations, the subsolar [$\\alpha$/Fe] ratios (Van Zee et al. 2004a) and the UV to near-IR spectral energy distribution of dE in the Virgo cluster (Gavazzi et al. 2002a), however, all indicate a more gradual star formation history with recent episodes of activity in these low-luminosity quiescent systems than in massive ellipticals. This constitutes the first evidence against their very old origin. A further disagreement with model predictions is that, although more frequent than bright galaxies, their number density is still significantly smaller than predicted by hierarchical models of galaxy formation for the field luminosity function (Kauffmann et al. 1993; Cole et al. 1994; Somerville \\& Primack 1999, Nagashima et al. 2005) or for the local group (Klypin et al. 1999; Bullock et al. 2000).\\\\ The strong similarities in the structural properties of star forming and quiescent dwarf galaxies in the nearby universe, namely their similar optical morphology and light profiles (both roughly exponentials), suggested that dwarf ellipticals might result from gas removal and subsequent suppression of star formation in gas rich dwarf galaxies. Gas removal might result either from its blowout due to the kinetic energy injected into the ISM by supernova explosion following a strong burst of star formation (Dekel \\& Silk 1986, Vader 1986, Yoshii \\& Arimoto 1987), gas exhaustion through subsequent episodes of star formation (Davies \\& Phillipps 1988), or from external perturbations induced by the hostile environment in which galaxies evolve. External perturbations include tidally induced mass loss in high speed encounters (Moore et al. 1998), tidal stirring (Mayer et al. 2001a, 2001b) and ram-pressure stripping induced by nearby companions (Lin \\& Faber 1983) or by the hot intergalactic medium in massive clusters (van Zee et al. 2004b). \\\\ Several observational evidences favor the environmental scenario against the gas blowout due to supernova explosions. The clearest indication that the environment plays an important role in the formation and evolution of dwarfs is the morphology segregation effect (Dressler 1980) which extends to low-luminosity systems (Binggeli et al. 1988; 1990; Ferguson \\& Binggeli 1994). Furthermore it has been recently claimed that the removal of the ISM through supernova winds is quite difficult in low-luminosity, dark matter dominated systems (Mac Low \\& Ferrara 1999; Ferrara \\& Tolstoy 2000; Silich \\& Tenorio-Tagle 1998, 2001). The discrete star formation history of several dwarf galaxies in the local group, as deduced by the analysis of their colour magnitude relation (Mateo 1998; Grebel 1999), is a further indication that dwarf spheroidals can retain their ISM through several episodes of star formation.\\\\ Another observational evidence favoring the transformation of star forming galaxies into quiescent dwarf ellipticals is the presence of rotationally supported (Pedraz et al. 2002; Geha et al. 2003; van Zee et al. 2004b) and/or HI gas rich (Conselice et al. 2003b; van Zee et al. 2004b) dE in the Virgo cluster. Recent studies based on SDSS imaging and spectroscopic data have shown that $\\sim$ 50 \\% of the bright end of the dE galaxy population in the Virgo cluster is characterized by disk features such as spiral arms or bars, this fraction decreasing down to $\\sim$ 5\\% at lower luminosities (Lisker et al. 2006a; see also Graham et al. 2003). Meanwhile $\\sim$ 15\\% of the bright dE in Virgo have blue centers revealing a recent activity of star formation. From a statistical point of view, the line-of-sight velocity distribution of dE inside clusters is similar to that of late-type galaxies suggesting a recent infall (Binggeli et al. 1993; Conselice et al. 2001).\\\\ Not all observational evidences, however, are consistent with the transformation of dwarf star forming galaxies into dwarf ellipticals under the effect of the environment. To reproduce the color magnitude relation of dwarf ellipticals, it has been claimed that magellanic irregulars should have faded $\\sim$ 1.5 magnitudes in the B band thus reaching surface brightnesses weaker than $\\mu(B)_e$ $=$ 25 mag arcsec$^{-2}$, values significantly smaller than the observed ones (Bothun et al. 1986). Most of the bright dwarf ellipticals in the Virgo cluster have a bright nucleus (Binggeli et al. 1985; Ferguson \\& Binggeli 1994) of small size, as shown by HST observations ($\\sim$ 4 pc, C\\^ote et al. 2006) while dwarf irregulars do not. While both dwarf irregulars and dwarf ellipticals follow different metallicity-luminosity relations, dwarf spheroidals are more metal-rich than dwarf irregulars at the same optical luminosity (Grebel et al. 2003 and references therein). The flattening distribution of non nucleated dE is similar to that of late-type spirals, Im and BCD (Binggeli \\& Popescu 1995). This class of objects, however, is significantly less round than nucleated systems. Although they exist, the fraction of objects belonging to the intermediate dIrr/dE transition class is too small. Furthermore a simple transformation of Im galaxies recently infalling in the cluster into dE does not seem to reproduce the observed difference in the cluster and field luminosity functions (Conselice 2002). Other strong constraints against the transformation of star forming into quiescent dwarfs are given by the studies of globular clusters, taken as a probe of the early phases of galaxy formation. The specific frequency of globular clusters in dwarf ellipticals, in fact, is significantly higher than that of star forming galaxies. This result has been interpreted as an evidence for a different formation scenario for the two galaxy populations (Miller et al. 1998; Strader et al. 2006).\\\\ With the aim of explaining these evidences, slightly different evolutionary scenarios have been proposed: cluster dwarf ellipticals might have been formed from higher surface brightness BCD galaxies (Bothun et al. 1986), or from tidally induced mass loss in multiple high speed encounters of massive galaxies (galaxy harassment; Mastropietro et al. 2005). Alternatively, Lisker et al. (2006a; 2006b; 2007) proposed that the cluster dwarf elliptical galaxy population is composed of different subcategories of objects, of which not all have been formed from gas stripped star forming dwarfs. As here emphasized the issue is still hotly debated.\\\\ A few years ago we started collecting multifrequency data covering the whole electromagnetic spectrum for galaxies in nearby clusters in order to study the effects of the environment on galaxy evolution. Up to now our research was primarily focused on the bright end of the luminosity function. Our interest covered the present and past star formation activity (Gavazzi et al. 1991, 1998, 2002b, 2006a), the atomic (Gavazzi et al. 2005b, 2006b) and molecular gas content (Boselli 1994; Boselli et al. 1994, 1997a, 2002), the radio continuum and IR properties of cluster galaxies (Gavazzi \\& Boselli 1999; Gavazzi et al. 1991). The results of our analysis, combined with those obtained by other teams, are summarized in a recent review article (Boselli \\& Gavazzi 2006).\\\\ The aim of the present paper is to extend the multifrequency analysis to the low-luminosity end of the luminosity function with the purpose of studying the possible transformation of star forming objects into dwarf ellipticals. This exercise is here done using a complete sample of galaxies in the Virgo cluster. The novelty of this work compared to previous investigations is twofold: the analyzed sample has an unprecedented multifrequency spectral coverage from the UV to near-IR imaging data. Furthermore this unique dataset is compared to the predictions of multizone chemo-spectrophotometric models of galaxy evolution here adapted to take into account the perturbations induced by the cluster environment. The type of perturbation simulated by the models are ram-pressure stripping (Gunn \\& Gott 1972) and starvation (Larson et al. 1980). The combination of multifrequency data with these models on the galaxies NGC 4569 (Boselli et al. 2006) and NGC 4438 (Boselli et al. 2005b) indeed indicated how powerful this method is for studying and constraining the evolution of cluster galaxies. \\\\ Our previous investigations have emphasized that the most important parameter governing the evolution is the total mass as traced by the near-IR H band luminosity (Gavazzi et al. 1996; 2002a; Boselli et al. 2001). Below a certain mass ($L_H$ $<$ 10$^{9.6}$ L$_{H \\odot}$) we identify the sequence of dwarf galaxies that we subdivide into quiescent and star forming disregarding their detailed morphology.\\\\ A major uncertainty in the interpretation of the multifrequency data, and in particular of those at short wavelength, is the extinction correction, which however is expected to be minor in low-metallicity star forming dwarf systems (Buat \\& Xu 1996) and probably negligible in dwarf ellipticals. For these reasons we decided to limit the present analysis to the cluster dwarf galaxy population, leaving the discussion relative to massive galaxies to a future communication. With the aim of driving the reader's attention to the scientific results of this work, the presentation of the dataset and the general description of the models are given in Appendix. ", "conclusions": "We can conclude that models and observations are consistent with an evolution of star forming, low-luminosity late type galaxies recently accreted in Virgo into quiescent dwarfs because of the ram pressure gas stripping and the subsequent stopping of their star formation activity. For consistency with surface brightness measurements, we show that high surface brightness star forming dwarf galaxies (low luminosity spirals and BCD) might be at the origin of the optically selected dwarf ellipticals (both normal and nucleated) analyzed in this work, although we expect that the low surface brightness magellanic irregulars (Sm and Im) once stripped of their gas reservoir, produce quiescent dwarfs with surface brightnesses below the detection limit of the VCC, as those observed in Virgo by Sabatini et al. (2005). The process of transformation is extremely rapid and efficient, since it works on all dwarfs and last on average less than 150 Myr. On longer time scales galaxies get structural and spectrophotometric properties similar to that of dwarf ellipticals. The whole star forming dwarf galaxy population dominating the faint end of the field luminosity function (Blanton et al. 2005), if accreted, can be totally transformed by the cluster environment into dwarf ellipticals on time scales as short as 2 Gyr and thus be at the origin of the morphology segregation observed also at low luminosities.\\\\ This interesting result is of fundamental importance even in a cosmological context because it shows that the majority of dwarf galaxies (if not all of them) are ``young'' even in clusters, and not old as expected in a hierarchical galaxy formation scenario (see also Nelan et al. 2005). For comparison with De Lucia et al (2006), we give in Table \\ref{Tabage} the lookback time when 50\\% and 80\\% of the stars were formed for a model galaxy of $V_C$ $=$ 55 km s$^{-1}$ for a (or two) ram pressure stripping event of efficiency $\\epsilon_0$ = 1.2 M$\\odot$ kpc$^{-2}$ yr$^{-1}$. \\noindent Although the values given in Table \\ref{Tabage} are not directly comparable to those given by De Lucia et al. (2006) since limited to relatively high mass objects (2.5 10$^9$ M$\\odot$, while our model galaxy is of only 2.4 10$^8$ M$\\odot$), the mean age of the stellar population of low luminosity, quiescent galaxies is significantly younger than that predicted by hierarchical models of galaxy formation (lookback times $\\sim$ 10 and 8.5 Gyr for 50\\% and 80\\% of the stellar population).\\\\ Despite their morphological type, the star formation activity of dwarf systems has been abruptly interrupted by the interaction with the environment in quiescent systems. This conclusion can be extended to the local group, where the study of the stellar color magnitude relation of dwarf spheroidal systems revealed that the star formation activity, although in an episodic manner, lasted for several Gyr (Mateo 1998; Grebel 1999). We can add that the only cluster galaxy population likely to be issued by major merging events, as those predicted by hierarchical models of galaxy formation, is that of massive ellipticals, whose origin is probably very remote ($z$ $\\geq$ 2-3, Dressler 2004; Treu 2004; Nolan 2004; Franx 2004; Renzini 2006). Indeed this is the only Virgo cluster galaxy population with a virialized velocity distribution (Conselice et al. 2001). If we consider clusters of galaxies as those regions where merging events were more frequent at early epochs since the big bang just because characterized by an high galaxy density, we can conclude that the hierarchical formation scenario was the principal driver of galaxy evolution only in massive objects at very early epochs. All observational evidences are consistent with a secular evolution afterward." }, "0801/0801.0266_arXiv.txt": { "abstract": "We study the alignment of grains subject to both radiative torques and pinwheel torques while accounting for thermal flipping of grains. By pinwheel torques we refer to all systematic torques that are fixed in grain body axes, including the radiative torques arising from scattering and absorption of isotropic radiation. We discuss new types of pinwheel torques, which are systematic torques arising from infrared emission and torques arising from the interaction of grains with ions and electrons in hot plasma. We show that both types of torques are long-lived, i.e. may exist longer than gaseous damping time. We compare these torques with the torques introduced by E. Purcell, namely, torques due to H$_2$ formation, the variation of accommodation coefficient for gaseous collisions and photoelectric emission. Furthermore, we revise the Lazarian \\& Draine model for grain thermal flipping. We calculate mean flipping timescale induced by Barnett and nuclear relaxation for both paramagnetic and superparamagnetic grains, in the presence of stochastic torques associated with pinwheel torques, e.g. the stochastic torques arising from H$_2$ formation, and gas bombardment. We show that the combined effect of internal relaxation and stochastic torques can result in fast flipping for sufficiently small grains and, because of this, they get thermally trapped, i.e. rotate thermally in spite of the presence of pinwheel torques. For sufficiently large grains, we show that the pinwheel torques can increase the degree of grain alignment achievable with the radiative torques by increasing the magnitude of the angular momentum of low attractor points and/or by driving grains to new high attractor points. ", "introduction": "Polarization of radiation arising from emission or absorption by aligned grains is widely used to study magnetic fields in the diffuse interstellar medium (see Goodman et al. 1995), molecular clouds (see Hildebrand et al. 2000, 2002) and in prestellar cores (Crutcher et al. 2004). The grain alignment involves the alignment of the axis of major inertia $\\ba_{1}$ with angular momentum $\\bJ$ and the alignment of $\\bJ$ with respect to the ambient magnetic field. An understanding of grain dynamics is extremely important for both understanding of dust properties and dust alignment. The latter is essential for determining situations when polarization of starlight passing through dusty interstellar gas, as well as polarization of dust emission can be reliably interpreted in terms of magnetic field direction. In terms of dust properties, fast rotation should disrupt loose aggregates, placing limits for the fractal dimensions of dust particles. As dust scatters or absorbs photons, it experiences radiative torques. These torques can be stochastic or systematic. Stochastic radiative torques arise from, for instance, a spheroidal grain randomly emitting or absorbing photons. The latter process, for instance, was invoked by Harwit (1970) in his model of grain alignment based on grains being preferentially spun up in the direction perpendicular to the photon beam. However, Purcell \\& Spitzer (1971) showed that randomization arising from the same grain emitting thermal photons makes the achievable degree of grain alignment negligible. In fact, they showed that random radiative torques arising from grain thermal emission is an important process of grain randomization. More recently, grain thermal emission was analyzed as a source of the excitation of grain rotation as well as its damping in relation to the rotation of tiny spinning grains that are likely to be responsible for the so-called anomalous foreground emission (Draine \\& Lazarian 1998). Systematic radiative torques were first introduced by Dolginov (1972) in terms of chiral, e.g. hypothetical quartz grains. Starlight passing through such grains would spin them up. Later, Dolginov \\& Mytrophanov (1976) considered an irregular grain model, consisting of two twisted spheroids, made of more accepted materials, e.g. silicate grains, and claimed that these grains will be both spun up and aligned by radiative torques arising from the anisotropic component of radiation field.\\footnote{Using so-called Rayleigh-Gans approximation, Dolginov \\& Silantiev (1976) provided calculations of the radiative torques for a model of an irregular grain consisting of two ellipsoids twisted with respect to each other, but our calculations of radiative torques using the Discrete Dipole Approximation code (Draine \\& Flatau 2004) for the same set parameters and the same model as in the aforementioned work are in conflict with their analytical findings (see more in Hoang \\& Lazarian 2008b).} Further on, following the convention we adopted in Lazarian \\& Hoang (2007a, hereafter LH07a), we shall call these torques RATs. The work by Dolginov \\& Mytraphanov (1976) was, unfortunately, mostly ignored for 20 years. A possible explanation of this may be due to the fact, that for a long time, the magnitude of RATs for realistic irregular grains remained unclear. A renewed interest to RATs was induced by the possibility of calculating them for arbitrary grain shapes. This occurred after Bruce Draine modified correspondingly his publicly-available Discrete Dipole Approximation code (hereafter DDSCAT; Draine \\& Flatau 2004). Moreover, Draine \\& Weingartner (1996, 1997, henceforth DW96, DW97, respectively) conjectured that RATs may provide the primary alignment mechanism for interstellar grains. Support for this claim came through later research in the field, by better understanding the dynamics of grains subject to RATs from anisotropic radiation fields, calculating grain alignment with an analytical model (LH07a), as well as including important physical processes like grain wobbling (Lazarian 1994; Lazarian \\& Roberge 1997; Weingartner \\& Draine 2003; Hoang \\& Lazarian 2008a), and accounting within the model for gaseous bombardment and uncompensated torques arising from H$_2$ formation, as was done in Hoang \\& Lazarian (2008a, henceforth HL08a). As it was described in a review by Lazarian (2007), RAT alignment mechanism has become not only the leading candidate to explain interstellar grain alignment, but also to explain grain alignment in many other astrophysical environments, including circumstellar regions (Aitken et al. 2002), accretion disks (Cho \\& Lazarian 2007), comet atmospheres (see Rosenbush et al 2007), and molecular clouds (Whittet et al. 2008). In LH07a, we subjected to scrutiny the properties of RATs. Using a simple analytical model (AMO) of a helical grain we studied the properties of RATs and the RAT alignment. The results obtained by the AMO were shown to be in good correspondence with numerical calculations obtained by DDSCAT for irregular grains. Invoking the generic properties of the RAT components, we explained the RAT alignment of grains in both the absence and presence of magnetic fields. Intentionally, for the sake of simplicity, in LH07a we studied a simplified dynamical model to demonstrate the effect of RATs. Within the latter model we showed that RATs can align grains at attractor points with low angular momentum (i.e., $J$ of the order of thermal angular momentum $J_{th}$, hereafter, low-$J$ attractor points), and/or attractor points with high angular momentum (i.e., $J\\gg J_{th}$, hereafter high-$J$ attractor points). The high-J attractor points mostly correspond to a perfect alignment of $\\bJ$ with respect to $\\bB$, while the low-J attractor points occur at perfect alignment angle or at some angle in the vicinity of the perfect alignment. One of the effects that was not discussed in LH07a was the effect of thermal fluctuations. Lazarian (1994) noticed that in spite of the fast rates of internal relaxation, rotating grains wobble. This conclusion was a consequence of the Fluctuation-Dissipation Theorem (see Landau \\& Lifshitz 1976), which states that any dissipation process, e.g. internal relaxation, should be accompanied by the proportionally fluctuation process. It is possible to show that grains rotating thermally are expected to wobble with larger amplitude compared to their counterparts rotating at suprathermal (much greater than thermal) rates (see Lazarian 1994, Lazarian \\& Roberge 1997). Therefore the zero value of angular momentum at low-J attractor points obtained in LH07a stemmed from the assumption of the perfect coupling of grain axis of the maximal moment of inertia $\\ba_{1}$ with $\\bJ$, which is violated for sufficiently low $J$. In fact, in HL08a, we treated fully the dynamics of the grain alignment by taking into account thermal fluctuations (see also Weingartner \\& Draine 2003), and showed that thermal fluctuations within irregular grains can increase the angular momentum of the low-J attractor points from $J=0$ to $J\\sim J_{d}$, i.e. the angular momentum of grains corresponding to dust grain temperature. In addition, HL08a showed that gas bombardment and other randomizing torques can contribute significantly to the RAT alignment by moving grains from low-$J$ to high-$J$ attractor points, whenever the latter are present. However, we feel that a comprehensive study on the degree of alignment as a function of grain size and radiation intensity is very essential for polarization modeling. Note that earlier papers which attempted to introduce grain alignment into polarization modeling (Cho \\& Lazarian 2005; Pelkonen et al. 2007; Bethell et al. 2007) assumed perfect alignment for grains larger than some critical size, and no alignment for smaller grains. The latter was inferred using the criterion of whether RATs for radiation field under interest, which are sharp function of grain size, can induce grains to rotate several times faster than the thermal rotation rate. This is a rather crude criterion because it does not take into account the type of RAT alignment, i.e with or without high-J attractor points. For instance, if RATs align all grains at low-$J$ attractor points, then assumption above leads to overestimates for the degree of alignment. In addition to RATs, other uncompensated torques act on grains (Purcell 1979; LD99a; Roberge \\& Ford 2000, henceforth RF00). Purcell (1979) proposed three processes that can produce uncompensated torques: hydrogen formation, photoelectric effect and the variation of accommodation coefficient. These pinwheel torques together with the paramagnetic dissipation were thought to be the major mechanism leading to the alignment of $\\bJ$ with $\\bB$. However, the paramagnetic dissipation was found to be unable to explain the alignment of paramagnetic grains in the weak interstellar magnetic field. In any case, in the presence of RATs, the alignment arising from paramagnetic dissipation for ordinary paramagnetic grains is negligible compared to that arising from RATs\\footnote{Unfortunately, the misconception that RATs act as proxies of Purcell's torques and the alignment arises from paramagnetic dissipation is well entrenched in the grain alignment literature. This claim is based on the DW96 study and, in spite of the fact, that the efficient RAT alignment was demonstrated already in DW97 and later works, the claim persisted. This may be due to the fact that, DW97 did not provide, unlike DW96, the predictions for the degree of alignment. The problem of predicting the degree of alignment happened to be a tough one. The present paper is an attempt in this direction. Interestingly enough, Dolginov \\& Mytrophanov (1976) did consider RATs as a mechanism capable of aligning grains irrespectively of the presence or absence of paramagnetic relaxation. For superparamagnetic grains with sufficient number of iron atom per cluster, Lazarian \\& Hoang (2008) showed that the alignment always happen with high-$J$ attractor points and is perfect. However, even in this case, RATs are more important than the enhanced paramagnetic relaxation. The latter effect causes the stabilization of the high-$J$ attractor point created by RATs.}. The efficiencies of the Purcell torques decrease if grains wobble thermally as we discussed above. The effect of thermal wobbling was quantified in Lazarian \\& Roberge (1997) and the results of this study were used by Lazarian \\& Draine (1999ab, henceforth LD99ab) to predict new effects of grain dynamics, namely, thermal flipping and thermal trapping. The thermal flipping is the effect of occasional increase of the wobbling angle beyond 90 degrees. When this occurs, the torques that are fixed in grain body coordinates change their direction. In the case thermal flipping occurs fast enough, the Purcell torques get averaged out and the grain rotates thermally in spite of the presence of the uncompensated pinwheel torques, i.e it is thermally trapped. However, this picture was challenged recently by Weingartner (2008). His study has several new mathematical points. The most important one is that he used a dimensionless variable $q=2I_{1}E/J^{2}$ with $I_{1}$ being the inertia moment along the axis of major inertia $\\ba_{1}$ and $E$ being the total rotational energy to describe the internal relaxation, instead of the angle $\\theta$ between $\\ba_{1}$ and $\\bJ$. He suggested an integration constant for diffusion coefficient of internal relaxation, which differs from that in the Lazarian \\& Roberge (1997). We agree with the choice of the integration constant. For this choice in Weingartner (2008) the diffusion coefficient vanishes when $\\ba_{1}$ perpendicular to angular momentum $\\bJ$. As a result, he found that grains do not experience thermal flipping as a result of internal relaxation. We agree with this choice of the integration constant as it corresponds to the nature of the Fluctuation-Dissipation Theorem employed in Lazarian (1994) to describe the wobbling of the grains. Indeed, according to the theorem no dissipation should correspond to no fluctuations. For an oblate grain the internal dissipation goes to zero as $\\theta \\rightarrow \\pi/2$. Thus the Fluctuation-Dissipation Theorem suggests that the at this point the fluctuations should also go to zero. One can check that the choice of the constant in Weingartner (2008), which was obtained from more formal considerations, corresponds to this physical requirement. However, as it clear from the rest of the paper, we disagree that this modification of the diffusion coefficients will preclude physical processes of thermal flipping and trapping introduced in Lazarian \\& Draine (1999) from happening. In what follows, we consider a more realistic picture of suprathermal rotation, namely, we take into account the fact that any pinwheel torque is accompanied by stochastic torques of the same nature. For instance, apart from systematic torques, the process of H$_2$ formation induces stochastic torques. In this model it is clear that the diffusion is present even when the angle between ${\\bf a_1}$ and ${\\bf J}$ is 90 degrees, which means that the thermal flipping is possible. In this paper we revisit the LD99a study and confirm that the thermal flipping and thermal trapping predicted in LD99a are important effects of grain dynamics. The structure of the present paper is as followings. In \\S 2 we present pinwheel torques, including long-lived and short-lived torques. We calculate the maximal value of angular momentum that these pinwheel torques can achieve for the ISM. In \\S 3 we revise the problem of thermal flipping and calculate the thermal mean flipping time induced by internal relaxation in the presence of external stochastic torques. Critical size of flipping grain and trapping size for the ISM are estimated in this section. We study the effects of H$_{2}$ pinwheel torques and thermal flipping by stochastic torques accompanied with H$_{2}$ formation on the RAT alignment for the diffuse ISM in \\S 4. We also calculate the value of $J$ at low attractor points as a function of the magnitude of H$_{2}$ pinwheel torques for different grain sizes in this section. We summarize our findings in \\S 5. ", "conclusions": "\\subsection{Grains are flipping and get trapped} New elements of grain dynamics, namely, thermal flipping and thermal trapping were reported in LD99a. The validity of these findings was challenged in Weingartner (2008), who found that grains can not experience thermal flipping as a result of internal relaxation. In the present work we showed that although internal relaxation can not result in grain thermal flipping, but in the presence of impulses due to H$_{2}$ formation and gas bombardment, grains can flip, and get trapped. The impulses play the role of an engine to bring grains through the boundary $\\ba_{1}\\perp\\bJ$. The thermal trapping is an essential process that, according to LD99a, explains why, in accordance with observations (see Kim \\& Martin 1995), small grains are not aligned by the Purcell (1979) paramagnetic relaxation process. The alternative explanations to this observational fact (see Lazarian 1995) are much less appealing. The sizes of the thermally trapped grains are not different from those which can be obtained using the LD99a treatment, but smaller than those discussed in LD99b. As a result, pinwheel torques are more important for interstellar grains than one could infer from LD99b. \\subsection{Extending theory of crossovers} Crossovers are times when grains subject to pinwheel torques slow down, then flip and get accelerating. The original theory of crossovers was suggested by Spitzer \\& McGlynn (1979). Lazarian \\& Draine (1997) improved the theory by accounting for the value of the residual thermal angular momentum, associated with thermal wobbling of the grain (see Lazarian 1994). LD99b showed that with nuclear relaxation taken into account the theory for regular crossovers in the spirit of the aforementioned works is applicable only for grains larger than $\\sim 10^{-4}$~cm, which excludes most of grains in diffuse interstellar gas. LD99a, instead, introduced the concepts of thermal flipping and trapping for smaller grains. The present paper shows that the flipping and trapped grains are smaller than those discussed in earlier papers. Therefore, in the absence of RATs, there is a range of grains for which the crossovers happen over the time scale larger than the internal relaxation time. In this situation the thermal value of the angular momentum can decrease below $J_d$ a result of the action of the pinwheel torques. This increases the randomization of grains during crossovers for the range of sizes larger than the flipping size. \\subsection{How common is suprathermal rotation?} Suprathermal, i.e. much faster than thermal, rotation was assumed to be default for most of the interstellar grains after the Purcell (1979) study. Studies of RATs in DW96, DW97 only made this point stronger. However, later works that reported thermal trapping of grains (LD99a, LD99b) and low-J attractor points (Weingartner \\& Draine 2003, LH07a) made one wonder whether most grains rotate thermally. The study in HL08a revealed that in the presence of gaseous bombardment or other randomization processes, the grains tend to diffuse from the low-J to high-J attractor point, provided that such high-J attractor point coexist with the low-J one. According to the parameter space study in LH07a (see Fig. 24 in LH07a) this corresponded to a fraction of grains that could vary depending on the angle between the radiation direction and the magnetic field, as well as on the $q^{max}$ ratio. In the present study we showed that for sufficiently large grains the pinwheel torques, provided that they are comparable with or stronger than the RATs, can induce suprathermal rotation of grains. Interestingly enough, this can happen even in the situations when only low-$J$ attractor points exist. Therefore, the presence of pinwheel torques can increase the percentage of the grains rotating suprathermally. However, it is worth noting, that for a substantial portion of the parameter space radiative torques act against pinwheel torques attempting to decrease the grain rotation rate. \\subsection{Implications to polarization modeling and magnetic field diagnostics} A quantitative theory of grain alignment is very important to polarization simulations. A number of works dealing with polarization simulations assume the perfect alignment of the grain axis with respect to the magnetic field (see CL07; Bethell et al. 2007; Pelkonen et al. 2007; Falceta-Goncalves, Lazarian \\& Kowal 2008). A detailed comparison between the polarization efficiency predicted by the theory of grain alignment and observations of dense cloud in Whittet et al. (2008) shows that a high degree of alignment is required to explain the observed polarization degree. How can this be accomplished? In Lazarian \\& Hoang (2008) we proposed one way of increasing the degree of grain alignment, namely, we noticed that in the situation when grains have superparamagnetic inclusions the high-$J$ attractor points are present for all the angles between the radiation direction and the magnetic field, as well as on the $q^{max}$ ratio. Thus the gaseous bombardment is expected to move all grains to the high attractor points, making the alignment perfect. In the present paper we show an alternative way of increasing the degree of grain alignment. In particular, we show that RATs plus pinwheel torques can enhance the degree of alignment by increasing the value of angular momentum at the low-$J$ attractor point. However, the circumstances for which the pinwheel torques affect the RAT alignment are rather restrictive: the pinwheel torques are required to be comparable or stronger than RATs. With these theoretical results one can attempt to distinguish between the two alternatives above and simultaneously get insight into the grain composition. First of all, more studies similar to those done by Whittet et al. (2008) are required to make sure that the high alignment is not a result of favorable illumination geometry. Alternatively, one may make measurements in situations where the illumination geometry is known. Comparing these results with the analytical and numerical predictions (see LH07a, HL08a) one can establish with higher certainty that the grain alignment is enhanced compared to what is expected from ordinary paramagnetic grains subject to RATs in the absence of the pinwheel torques. In parallel, one can search for the correlation of grain alignment enhancement with the environments where we expect the higher amplitudes of the pinwheel torques. If such an enhancement is present, it will indicate that grains are {\\it not} superparamagnetic. If they were superparamagnetic, we would expect the perfect alignment for typical conditions of the diffuse ISM, which one cannot be improved further by the action of the pinwheel torques. No correlation, but high degrees of alignment mean, on the contrary, the existence of superparamagnetic grains. One can conclude by observing an interesting cycle in the development of the grain alignment theory. The initial theory by Davis-Greenstein (1951) appealed to paramagnetic dissipation in ordinary paramagnetic grains. The mechanism was shown not to be sufficiently strong, however. Then to enhance the efficiency of the initial Davis-Greenstein process both superparamagnetism (Jones \\& Spitzer 1967) and pinwheel torques (Purcell 1979) were appealed to. At the moment we may face the situation that the RATs are not strong enough to explain the alignment and again we appeal to superparamagnetism (Lazarian \\& Hoang 2008) and pinwheel torques (this paper). Nevertheless, there is a substantial difference between the two situations. The paramagnetic relaxation in the diffuse interstellar medium could provide a few percent alignment at most. The RAT alignment provides the degrees of alignment about 20\\% even in the cases of unfavorable directions of illumination (see HL08a). In addition, while the Davis-Greenstein process favored the alignment of small grains, the RAT alignment, in accordance with observations, favors the alignment of large grains. \\subsection{Summary} Our main results are summarized as follows: \\begin{itemize} \\item We extended the definition of the pinwheel torques to include three new processes, namely, radiative torques arising from grain infrared emission, torques due to interactions of grain with electrons in plasma, as well as, torques arising from gas flow interacting with a helical grain. We proposed to consider the radiative torques arising from the action of isotroptic radiation on a grain as pinwheel torques. All these torques we identified as long-lived, as their life-time is expected to be longer than the typical grain rotational damping time. We used H$_2$ torques as a proxy of for general pinwheel torques, but, unlike earlier studies, we allowed the pinwheel torques to be both short-lived and long-lived. We found that in the ISM, the pinwheel torques due to infrared emission is in general weaker than that due to H$_{2}$ formation, and RATs. \\item We proved that flipping and trapping of astrophysical dust grains is the active and important process. We did this by augmenting the torques arising from internal relaxation by the stochastic torques that inevitably accompany the action of pinwheel torques and gas bombardment. . We calculated the thermal flipping induced by the Barnett, nuclear relaxations, as well, as internal relaxation in superparamagnetic grains in the presence of impulses due to H$_{2}$ formation and gas bombardment. Adopting the revised diffusion coefficients for internal relaxation from Weingartner (2008), we found that flipping is fast for small grains, and increases when superparamagnetic inclusions are accounted for. We obtained the critical size of flipping and trapping size of the thermally trapped grains for the ISM. \\item The most important result of this study is related to the increase of the expected degree of grain alignment when dust grains are subject to both RATs and pinwheel torques. This increase stems from the increase by pinwheel torques of the value of the angular momentum at the low-$J$ attractor point. The increase of angular momentum decreases the wobbling of grains at the low-$J$ attractor point and therefore increase the degree of internal alignment as the internal alignment (of $\\ba_{1}$ with $\\bJ$) is nearly perfect when the grain rotates with rates much larger than the thermal rotation rate. The joint pinwheel plus RATs alignment is different from the Purcell (1979) alignment. In the latter mechanism the alignment is due to paramagnetic dissipation, which induces torques that are very weak compared to RATs. \\end{itemize}" }, "0801/0801.2749_arXiv.txt": { "abstract": "We present Very Long Baseline Array (VLBA) observations of the TeV blazars H\\,1426+428, 1ES\\,1959+650, and PKS\\,2155$-$304 obtained during the years 2001 through 2004. We observed H\\,1426+428 at four epochs at 8\\,GHz, and found that its parsec-scale structure consisted of a $\\sim 17$\\,mJy core and a single $\\sim 3$\\,mJy jet component with an apparent speed of $2.09c\\pm0.53c$. The blazar 1ES\\,1959+650 was observed at three epochs at frequencies of 15 and 22\\,GHz. Spectral index information from these dual-frequency observations was used to definitively identify the core of the parsec-scale structure. PKS\\,2155$-$304 was observed at a single epoch at 15\\,GHz with dual-circular polarization, and we present the first VLBI polarimetry image of this source. For 1ES\\,1959+650 and PKS\\,2155$-$304, the current observations are combined with the VLBA observations from our earlier paper to yield improved apparent speed measurements for these sources with greatly reduced measurement errors. The new apparent speed measured for component C2 in 1ES\\,1959+650 is $0.00c\\pm0.04c$ (stationary), and the new apparent speed measured for component C1 in PKS\\,2155$-$304 is $0.93c\\pm0.31c$. We combine the new apparent speed measurements from this paper with the apparent speeds measured in TeV blazar jets from our earlier papers to form a current set of apparent speed measurements in TeV HBLs. The mean peak apparent pattern speed in the jets of the TeV HBLs is about 1$c$. We conclude the paper with a detailed discussion of the interpretation of the collected VLBA data on TeV blazars in the context of current theoretical models for the parsec-scale structure of TeV blazar jets. ", "introduction": "\\label{intro} The field of TeV gamma-ray astronomy has grown rapidly over the past several years with the advent of sensitive new gamma-ray telescopes such as H.E.S.S.\\ and MAGIC. Particularly interesting results have been obtained in blazar astronomy, as the total number of detected TeV blazars has increased from six only a few years ago to 17 this year (Wagner 2007b). The majority of the detected TeV blazars (16 of 17) belong to the class of high-frequency peaked BL Lac objects, or HBLs --- so named because the two peaks in their spectral energy distribution (SED) occur at relatively high UV/X-ray and GeV/TeV energies. This two-peaked spectral energy distribution in HBLs is most commonly interpreted as the result of relativistic electrons (and possibly positrons) radiating in a jet which is undergoing bulk relativistic motion at a small angle to the observer's line of sight, with the low-frequency peak due to synchrotron radiation, and the high-frequency peak due to inverse-Compton scattering of the jet's own synchrotron photons (synchrotron self-Compton, or SSC emission). The TeV source list is mostly limited to relatively nearby blazars ($z\\lesssim0.2$), because of the absorption of TeV gamma-rays on the extragalactic background light. The TeV observations have shown dramatic variability in a number of these blazars. Among the most remarkable variability events are the 200\\,s variability timescale detected for PKS 2155$-$304 in July 2006 by H.E.S.S.\\ (Aharonian et al.\\ 2007), and the 3-minute variations detected for Mkn\\,501 in June 2005 by MAGIC (Wagner 2007a). Such rapid variations suggest extremely small emitting volumes and/or time compression by large relativistic Doppler factors, e.g., $\\delta\\gtrsim100$ for PKS 2155$-$304 (Aharonian et al.\\ 2007). High Doppler factors are also sometimes invoked in specific SSC models of TeV blazar spectra and variability; for example, Fossati et al. (2007) consider two SSC models to explain the X-ray/TeV variability of Mkn 421: one with $\\delta\\sim20$ (scattering in the Klein-Nishina regime), and one with $\\delta\\sim100$ (scattering in the Thomson regime). Such high Doppler factors are at the upper limit of what is expected in relativistic jets, and challenge our understanding of these objects if they do in fact occur. Complementary observations that are crucial to unraveling the physics of TeV blazar jets are provided by the VLBI technique, which yields radio images of the relativistic jets with sub-parsec resolutions for these nearby blazars. Apparent jet speeds, brightness temperatures, and limits on jet/counterjet brightness ratios can all be measured from VLBI images, and these quantities all provide constraints on fundamental jet parameters such as the bulk Lorentz factor and the angle of the jet to the line-of-sight (subject to some caveats discussed at length in $\\S$~\\ref{disc}). Our understanding of the parsec-scale radio properties of the HBL class in general has been increased by the recent work of Giroletti et al.\\ (2004a, 2006). Those authors studied a sample of low-redshift BL Lac objects with a variety of radio instruments, and demonstrated that the radio properties of the HBLs are consistent with them being the beamed versions of nearby low-luminosity FR I radio galaxies. The TeV blazars thus likely have an intrinsically different parent population with weaker jets, compared to the more distant powerful blazars, which likely have FR II parents (e.g., Urry \\& Padovani 1995). The mean derived parsec-scale Lorentz factor for the HBL class by Giroletti et al.\\ (2004a), including TeV sources, is only $<\\Gamma>\\sim3$, much lower than the Lorentz factors suggested by the TeV gamma-ray emission. We have previously published a series of papers investigating the parsec-scale kinematic properties of TeV-detected HBLs through multi-epoch high-resolution VLBI observations, predominantly with the National Radio Astronomy Observatory's Very Long Baseline Array (VLBA) \\footnote{The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.}, and we extend that study in this paper. Previous papers in this series have included: early VLBA and space VLBI observations of Mkn\\,421 (Piner et al.\\ 1999), VLBI observations of Mkn\\,501 (Edwards \\& Piner 2002), VLBA observations of 1ES\\,1959+650, PKS\\,2155$-$304, and 1ES\\,2344+514 (Piner \\& Edwards 2004), and new VLBA polarimetry observations of Mkn\\,421 (Piner \\& Edwards 2005). This fifth paper in the series adds new VLBA observations of the TeV blazars \\object{1ES 1959+650} and \\object{PKS 2155-304} whose kinematics were relatively poorly determined in Piner \\& Edwards (2004) (hereafter Paper~I), as well as a four-epoch series of VLBA observations of \\object{H 1426+428} --- the sixth TeV blazar to be discovered. Altogether we present six new images of 1ES\\,1959+650, four new images of H\\,1426+428, and one new polarization image of PKS\\,2155$-$304, for a total of eleven new datasets. These datasets comprise the observations from our TeV blazar monitoring program over the years 2001-2004 (excepting the observations of Mkn\\,421, which were published separately in Piner \\& Edwards [2005]). Note that except for the two brightest sources (Mkn\\,421 and Mkn\\,501), the TeV blazars are too faint in the radio ($\\lesssim$100\\,mJy) to be included in other VLBA monitoring programs such as MOJAVE, and they typically require long observations to obtain images of sufficient dynamic range. Observational backgrounds on the three specific sources studied in this paper are presented in the results section for the specific source. Our earlier work on the parsec-scale structure of TeV blazars has shown a noticeable lack of superluminal components in their jets, which contrasts with the rapid superluminal motions observed in the jets of more powerful blazars, and with the high Lorentz factors derived from modeling the TeV emission. We have previously interpreted the general lack of superluminal components in HBLs as evidence for a lower bulk Lorentz factor in the parsec-scale radio-emitting region compared to the TeV-emitting region. Models that have been invoked to explain this `bulk Lorentz factor crisis' include: jets that are decelerated along their length (Georganopoulos \\& Kazanas 2003; Wang, Li, \\& Xue 2004; Bicknell et al.\\ 2005), jets with transverse velocity structures consisting of a fast spine and a slower layer (Giroletti et al.\\ 2004b; Ghisellini, Tavecchio, \\& Chiaberge 2005; Henri \\& Saug\\'{e} 2006), or jets with opening angles large enough that unintentional averaging over multiple viewing angles (because of limited resolution) causes the apparent conflict (Gopal-Krishna, Dhurde, \\& Wiita 2004; Gopal-Krishna, Wiita, \\& Dhurde 2006). Finally, Gopal-Krishna et al.\\ (2007) consider a combination of the last two models; i.e., large opening angle jets with transverse velocity structures. Some of these models may be distinguished through the observed statistical distribution of apparent speeds in TeV blazar jets (e.g., Gopal-Krishna et al.\\ 2006), so we conclude the paper by presenting our current best set of TeV blazar apparent speed measurements from the complete series of five papers, which consists of sixteen component speeds in six sources, and discussing the various models in the context of the current observations. In this paper we use cosmological parameters $H_{0}=71$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{m}=0.27$, and $\\Omega_{\\Lambda}=0.73$ (Bennett et al.\\ 2003). When results from other papers are quoted, they have been converted to this cosmology. ", "conclusions": "We have presented new multi-epoch VLBI images for the three TeV blazars H~1426+428, 1ES~1959+650, and PKS~2155$-$304 obtained during the years 2001 to 2004. The results for H~1426+428 are the first multi-epoch VLBI results to be presented for this source. The results for 1ES~1959+650 and PKS~2155$-$304 are combined with earlier results for these sources from Paper I to yield a longer time baseline for measurement. The major observational findings for these three sources are: \\begin{enumerate} \\item{H~1426+428's parsec-scale structure during this time range was well-modeled by a $\\sim 17$\\,mJy core and a single $\\sim 3$\\,mJy jet component at a position angle of approximately $-25\\arcdeg$, with an apparent speed of $2.09c\\pm0.53c$.} \\item{1ES~1959+650 consisted of a compact core and a nearby stationary ($0.00c\\pm0.04c$) jet component at a position angle of about $125\\arcdeg$ and a separation of about 0.35 mas at 15 GHz. On larger scales of a few mas the jet is diffuse and directed to the north, so that this source shows an extreme apparent misalignment of about $130\\arcdeg$ on parsec scales.} \\item{PKS~2155$-$304 was observed with dual-circular polarization at 15 GHz. The fractional polarization at the position of the core was 3\\%, and the electric vector position angle was $131\\arcdeg$, about $30\\arcdeg$ misaligned from the innermost jet position angle. The measured apparent speed of the single jet component was $0.93c\\pm0.31c$.} \\end{enumerate} We combined the new apparent speed measurements from this paper with the apparent speeds measured in TeV blazar jets from our earlier papers to form a current set of apparent speed measurements in TeV HBLs (Table~\\ref{speedtab}). The mean peak apparent pattern speed in the jets of the TeV HBLs is about 1$c$. The statistical result noted in Paper I that the TeV HBLs have significantly slower apparent jet pattern speeds compared to radio-selected or GeV-selected blazars is strengthened by the new results of this paper. The Discussion section ($\\S$~\\ref{disc}) presented a thorough analysis of these results in the context of other radio observations and theoretical models for TeV blazar jets. Conclusions from that analysis were: \\begin{enumerate} \\item{The peak apparent speeds of order 0.5-2$c$ in four of the six studied TeV blazars (Table~\\ref{speedtab}), when taken together with other observed radio properties discussed in $\\S$~\\ref{disc}, are consistent with radio jets with bulk Lorentz factors of $\\Gamma\\sim3$ and viewing angles of a few degrees.} \\item{The very slow peak apparent speeds in two of the six studied TeV blazars in Table~\\ref{speedtab} (Mkn~421 and 1ES\\,1959+650) are likely pattern speeds that are unrelated to the bulk apparent speeds of the jets, which are likely to be similar to those mentioned above.} \\end{enumerate}" }, "0801/0801.2864_arXiv.txt": { "abstract": "{The Australian SKA Pathfinder (ASKAP) will be a powerful instrument for performing large-scale surveys of galaxies. Its frequency range and large field of view makes it especially useful for an all-sky survey of Local Volume galaxies, and will probably increase the number of known galaxies closer than 10 Mpc by a factor of two and increase, by at least an order of magnitude, the number detected in HI. Implications for our knowledge of the HI mass function for the very faintest galaxies and for the structure and dynamics of the Local Volume are discussed. } ", "introduction": "\\label{sec:intro} The Local Volume is a key region for the study of the properties of galaxies, including: (1) their internal structure and dynamics; (2) their spatial distribution and dynamics in an environment which lies in the outskirts of a supercluster; and (3) their complete evolutionary history, by virtue of our ability to resolve individual stars. The Local Volume is particularly useful for studying the faintest galaxies and is well-served by having a wealth of accurate redshift-independent TRGB distances through recent surveys by the HST \\cite{Riz07}. As shown elsewhere in these proceedings, detailed studies of Local Volume galaxies in the 21cm line of neutral hydrogen have been particularly fruitful. These have recently been rejuvenated by surveys at other wavebands including those of {\\em Spitzer} and {\\em GALEX}. Due to the sensitivity and resolution, such surveys have often been drawn towards the luminous galaxy population. However, the Local Volume also offers a unique opportunity to study the faintest galaxies observable and many studies (e.g. {\\em SINGG, LVHIS, THINGS}) have also been careful to select their samples across a range of intrinsic luminosity. The Square Kilometre Array (SKA) will be a radio telescope of unprecedented power to observe galaxies in the radio continuum and in the 21cm line. Its sensitivity will easily surpass existing telescopes and allow Local Volume galaxies to be studied at the highest spatial resolution. However, it's not due to come on-line for at least a decade. Nevertheless, the next generation of so-called `SKA pathfinders' are around the corner. Their purpose is to test SKA technologies, yet provide sufficient sensitivity to obtain useful science and provide a valuable source of survey material for the SKA itself. Examples of proposed pathfinders are the Allen Telescope Array (ATA) and the Apertif upgrade to WSRT in the northern hemisphere and MeerKAT and the Australian SKA Pathfinder (ASKAP) in the southern hemisphere. In this brief paper, I will look at the implications for our knowledge of the Local Volume of proposed ASKAP surveys. ", "conclusions": "The high number of galaxies that, in all likelihood, remain to be discovered in the Local Volume will allow a remarkably dense sampling of the extragalactic environment of the Local Group. This will allow an accurate mapping of the large-scale filamentary features joining the Local Group with Sculptor and other groups. Combined with the redshift-independent distances that are possible for such nearby objects, it will also allow a study of the Hubble flow, infall towards filaments, and tidal stretching owing to nearby overdense regions such as the Local Supercluster and underdense regions such as the Local Void." }, "0801/0801.0050_arXiv.txt": { "abstract": "In this talk we introduce our recent results of global 1D MHD simulations for the acceleration of solar and stellar winds. We impose transverse photospheric motions corresponding to the granulations, which generate outgoing Alfv\\'{e}n waves. The \\Alfven waves effectively dissipate by 3-wave coupling and direct mode conversion to compressive waves in density-stratified atmosphere. We show that the coronal heating and the solar wind acceleration in the open magnetic field regions are natural consequence of the footpoint fluctuations of the magnetic fields at the surface (photosphere). We also discuss winds from red giant stars driven by \\Alfven waves, focusing on different aspects from the solar wind. We show that red giants wind are highly structured with intermittent magnetized hot bubbles embedded in cool chromospheric material. ", "introduction": "The \\Alfven wave, generated by the granulations or other surface activities, is a promising candidate operating in the heating and acceleration of solar winds from coronal holes. It can travel a long distance so that the dissipation plays a role in the heating of the solar wind plasma as well as the lower coronal plasma, in contrast to other processes, such as magnetic reconnection events and compressive waves, the heating of which probably concentrates at lower altitude. While high-frequency ioncyclotron waves are recently highlighted for the preferential heating of minor heavy ions (Axford \\& McKenzie 1997) the protons, which compose the main part of the plasma, are supposed to be mainly heated by low-frequency ($\\lesssim 0.1$Hz) components in the MHD regime. because (1) the low-frequency wave is expected to have more power, and (2) the resonance frequency of the proton is higher than those of heavier ions so that the energy of the ioncyclotron wave is in advance absorbed by heavy ions (Cranmer 2000). In this paper, we focus on roles of such low frequency \\Alfven waves. When considering low-frequency \\Alfven waves in solar and stellar atmospheres, the stratification of the density due to the gravity is quite important because the variation scale of the density, and accordingly \\Alfven speed, is comparable to or shorter than the wavelengths; the WKB approximation is no longer applicable. Also, the amplitude is amplified because of the decrease of the density so that waves easily become nonlinear. Recently, we have extensively studied the heating and acceleration of solar and stellar winds by self-consistent MHD simulations from the photosphere to sufficiently outer region (Suzuki \\& Inutsuka 2005; 2006; hereafter SI06; Suzuki 2007). We review these works in this contribution talk. ", "conclusions": "We have performed 1D MHD numerical simulations of solar and stellar winds from the photosphere. The low-frequency \\Alfven waves are generated by the footpoint fluctuations of the magnetic field lines. We have treated the wave propagation and dissipation, and the heating and acceleration of the plasma in a self-consistent manner. Our simulation is the first simulation which treats the wind from the real surface (photosphere) to the (inner) heliosphere with the relevant physical processes. We have shown that the dissipation of the low-frequency \\Alfven waves through the generation of the compressive waves (decay instability) and shocks (nonlinear steepening) is one of the solutions for the heating and acceleration of the plasma in the coronal holes. However, we should cautiously examine the validity of the 1-D MHD approximation we have adopted. There are other dissipation mechanisms due to the multidimensionality, such as turbulent cascade into the transverse direction (Goldreich \\& Sridhar 1995; Oughton et al.2001) and phase mixing (Heyvaerts \\& Priest 1983). % If \\Alfven waves cascade to higher frequency, kinetic effects (e.g. Nariyuki \\& Hada 2006) becomes important. We have also extended the solar wind simulations to red giant winds. With stellar evolution, the steady hot corona with temperature, $T \\approx 10^6$ K, suddenly disappears because the surface gravity becomes small; hot plasma cannot be confined by the gravity. Thermal instability also generate intermittent magnetized hot bubbles in cool chromospheric winds." }, "0801/0801.3187_arXiv.txt": { "abstract": "Recent observations of the solar corona with the LASCO coronagraph on board of the SOHO spacecraft have revealed the occurrence of triple helmet streamers even during solar minimum, which occasionally go unstable and give rise to large coronal mass ejections. There are also indications that the slow solar wind is either a combination of a quasi-stationary flow and a highly fluctuating component or may even be caused completely by many small eruptions or instabilities. As a first step we recently presented an analytical method to calculate simple two-dimensional stationary models of triple helmet streamer configurations. In the present contribution we use the equations of time-dependent resistive magnetohydrodynamics to investigate the stability and the dynamical behaviour of these configurations. We particularly focus on the possible differences between the dynamics of single isolated streamers and triple streamers and on the way in which magnetic reconnection initiates both small scale and large scale dynamical behaviour of the streamers. Our results indicate that small eruptions at the helmet streamer cusp may incessantly accelerate small amounts of plasma without significant changes of the equilibrium configuration and might thus contribute to the non-stationary slow solar wind. On larger time and length scales, large coronal eruptions can occur as a consequence of large scale magnetic reconnection events inside the streamer configuration. Our results also show that triple streamers are usually more stable than a single streamer. ", "introduction": "Recent observations of the corona with the LASCO coronagraph \\cite{schwenn97} on board of the SOHO spacecraft showed that the corona can be highly structured even during the solar activity minimum. The observations revealed a triple structure of the streamer belt which was existent for several consecutive days. The observations further showed that these triple structures occasionally go unstable leading to a seemingly new and extraordinarily huge kind of coronal mass ejection (global CMEs). Natural questions arising from these observations are whether the helmet streamer triple structure is directly connected with or responsible for the occurrence of global CMEs and what is the physical mechanism of their formation. The structure of helmet streamers and their stability has been studied both observationally and theoretically for a long time (e.g. \\opencite{pneuman:kopp71}; \\opencite{cuperman:etal90}; \\opencite{cuperman:etal92}; \\opencite{koutchmy:livshits92}; \\opencite{wang:etal93}; \\opencite{cuperman:etal95}; \\opencite{wu:etal95}; \\opencite{bavassano:etal97}; \\opencite{noci}; \\opencite{hundhausen99}). There seems to be a natural association of helmet streamers with coronal eruptions and coronal streamers are assumed to be the source region of the slow solar wind. The traditional view towards the origin of the slow solar wind is that it is a more or less stationary plasma flow on open field lines around the closed field lines of a helmet streamer. Recent observations (e.g. \\opencite{habbal98}; \\opencite{noci}) challenge this traditional view and indicate that the slow solar wind is non-stationary and seems to be produced and accelerated by small eruptions in the helmet streamer stalk above the cusp. This acceleration process of the slow solar wind has been compared with the rise of smoke above a burning candle (Schwenn, private communication). Pre-SOHO observations of multiple streamer configurations during the maximum phase of solar activity and of the multiple current sheet structure of the heliospheric plasma sheet \\cite{crooker:etal93,woo:etal95} have initiated several studies of the dynamics and stability of multiple current sheets with variations in only one spatial dimension \\cite{otto:birk92,yan:etal94, dahlburg:karpen95,birk:etal97,wang:etal97}. \\inlinecite{einaudi:etal99} have recently presented a model for the generation of the non-stationary slow solar wind based on linear and non-linear stability calculations for a single one-dimensional current sheet with field-aligned flow. All these models do, however, in a strict sense only apply to the streamer stalk, i.e.\\ to the open field lines of the heliospheric current sheet. Here we aim to investigate both closed and parts of the adjacent open field line regions. Models of multiple arcade and loop structures have been investigated before by e.g.\\ \\inlinecite{mikic:etal88}, \\inlinecite{biskamp:welter89} and most recently by \\inlinecite{antiochos:etal99}. Our model differs from these models basically by the possibility of having a flexible analytical initial condition for the time-dependent calculations allowing the investigation of different types of structures. As a first step towards improving the theoretical understanding of the above mentioned phenomena in triple streamer configurations, we have calculated analytic two-dimensional static models of triple helmet streamers in a previous paper (\\opencite{paper1}; further cited as {\\it paper I}). The aim of the present paper is to undertake the next step in this investigation and to study the stability of the stationary state helmet streamer configurations calculated in {\\it paper I}. We will carry this out with the help of time-dependent numerical experiments using the equations of resistive magnetohydrodynamics. The outline of the paper is as follows. In Section \\ref{basics} we discuss the basic equations and briefly describe our numerical method. Section \\ref{model} outlines our main model assumptions. In Section \\ref{results} we present the results of numerical experiments and in Section \\ref{conclusions} we discuss our results and give an outlook on future work. ", "conclusions": "\\label{conclusions} In this paper we have tried to make a step towards a better theoretical understanding of the dynamics of helmet streamers with triple structure. We investigated the possible role of triple streamers for the development of coronal mass ejections and as a possible source for plasma and magnetic field for the wind emanating from streamer regions. In previous works (e.g. \\opencite{steinolfson94}; \\opencite{linker94}; \\opencite{wu:etal95}; \\opencite{wu97} ) only single helmet streamer where modelled and these models assumed the slow solar wind as a stationary plasma flow on open field lines in the streamer region. To get a start equilibrium these authors solved the ideal time-dependent MHD equations numerically until a stationary state was reached. These works showed that helmet streamers can become unstable and produce coronal mass ejections. The present investigations where motivated by the new observations with the LASCO coronagraph on SOHO \\cite{schwenn97}. These observations showed, that the streamer belt in the solar activity minimum typically has a triple structure. The observations also gave further strong evidence that a stationary slow solar wind may not exist but is produced by many small eruptions. Apart from these continuously occuring small eruptions, also large, however rarely occuring coronal mass ejections are generated in the triple streamer belt. To take these observations into account in a helmet streamer model, we developed an analytical stationary model of triple helmet streamers using the ideal MHD equations in {\\it paper I}. The initial states have to possess a non-vanishing free energy to allow their instability with respect to magnetic reconnection. As discussed in {\\it paper I} we took into account the observation that the streamer configurations are very extended in the radial direction to simplify the calculation of the initial states. We emphasize that such radially extended configurations cannot be modeled by potential fields. In the present paper we investigated the stability of these stationary state configurations with the help numerical experiments in the framework of time-dependent resistive MHD. We used three ad hoc models for the resistivity, a constant resistivity, a resistivity localized at the thin current sheets and a current-dependent resistivity. We first investigated the ideal stability of our triple streamer configurations and found that they are stable on the time-scale of our simulations. Next we investigated the resistive stability of a triple streamer configuration without cusp structure. We found that our triple streamer configuration is resistively unstable. Reconnection takes place and plasmoids form in each of the three closed field line regions. By comparing the time evolution of the triple streamer model with a single streamer model we found that for a triple streamer configuration without cusp structure the time-evolution is usually slower than the corresponding time evolution of a single streamer. The triple streamer evolution also shows characteristic differences in comparison with the single streamer case concerning the location of the reconnection sites. We could explain these differences by the influence the three streamers exert on each other. For triple streamer configurations with cusp structure we found quite similar results for reconnection processes inside the closed field line regions but in addition we found that the helmet streamer stalk above the cusp is highly unstable to reconnection. This reconnection process leads to the formation of a dome above the triple structure, i.e. a region of closed field lines which encloses the triple structure completely. The resistive instability of the streamer stalk current sheet could be a possible source of plasma and magnetic field for the non-steady solar wind emanating from the streamer regions. In the present paper we have only been able to demonstrate that this mechanism works with a preexisting cusp structure. In later stages of our simulations the cusp is replaced by an X-point at which reconnection can take place. The inclusion of flow on open field lines would also allow for the possibility to generate a new cusp structure making a repetition of the process possible. Furthermore we found interaction between the two outer streamers just below the cusp region. This interaction can also contribute to the formation of the dome. This dome makes it more difficult for plasmoids to escape and thus streamers with cusp structure within our model are less likely to eject material than the configurations without cusp. These results are consistent with the observational finding that the triple streamer configuration is observed to be stable for several days. One may also speculate about the fact that the observations usually show three streamers which approximately have the same radial extension. A possible explanation on the basis of our model is that if one of the streamers grows and becomes much larger than the other streamers, it becomes prone to instability and a coronal mass ejection occurs similarly to the case of a single streamer. In this process the streamer looses energy, mass and magnetic flux and returns to its original state. The numerical experiments presented here can only be considered as a very first step towards a complete model of these interesting phenomena. We already mentioned above that although the models with cusp structure seem to be matching the observed streamer structure best of all our models, it is not possible to include regions of open field lines between the streamers in our models. One way to overcome this shortcoming would be to use three-dimensional models, which is a natural next step. Other possible improvements of the present work are the inclusion of gravity and the use of spherical geometry." }, "0801/0801.1527_arXiv.txt": { "abstract": " ", "introduction": "The standard model (SM) provides an accurate description of particle physics below the electroweak scale but it is generally though not to be valid up to arbitrary energies. Extensions of this celebrated scheme, invoked to cure diseases like the strong CP, hierarchy or flavor problems, often involve higher gauge symmetries and further matter content. Moreover, string theory is a preferred candidate for the unification of quantum mechanics and general relativity where additional gauge and matter fields are assured. At low energies, some of these new fields can arrange into a ``hidden sector\" if only very massive particles (or gravity) mediate interactions between them and the SM ``visible sector\". Of course, depending on the scalar content of the theory, gauge symmetries can either be spontaneously broken or remain exact. Then, the corresponding ``hidden\" bosons could have in principle an arbitrary mass. If this is small enough, these hidden bosons can have a very rich phenomenology at present affordable energy scales. The simplest case concerns just a novel $U(1)_\\mathrm{h}$ symmetry and its corresponding gauge boson, henceforth called ``hidden\" photon. The interplay between this hidden photon and the SM photon modifies the predictions of quantum electrodynamics \\cite{Okun:1982xi}, often claimed to be the most accurate of all physical theories so far, thus constraining the hidden photon parameters. We can turn this argument in the opposite direction: the constraints on hidden photon parameters give us information about how accurate is the QED description of nature at low energies. A number of laboratory experiments has been devoted to the search of hidden photons, the resulting bounds being strongly dependent on the hidden photon mass. For masses corresponding to macroscopic length scales, experiments testing the Coulomb law \\cite{Williams:1971ms,Bartlett:1988yy} set strong constraints on hidden photons, but still they could be largely improved by experiments dealing with high quality microwave cavities \\cite{Jaeckel:2007ch}. In the microscopic range, laser experiments are also becoming very powerful probes of hidden sector particles \\cite{Cameron:1993mr, Gies:2006ca,Ahlers:2006iz,Ahlers:2007rd,Jaeckel:2007gk,Robilliard:2007bq,Chou:2007zz,Ahlers:2007qf}. At atomic distances, comparison of the Rydberg constant for different atomic levels gives interesting but weak bounds \\cite{Beausoleil:1987,Garreau:1987}. Finally, particle colliders extend the mass range until typical electroweak scales \\cite{Kors:2004dx,Feldman:2006ce,Chang:2006fp,Kumar:2006gm,Feldman:2007nf}. On top of that, the evolution of stars turns out to be the most sensitive ``laboratory\" to study properties of novel low mass weakly interacting particles \\cite{Raffelt:1996wa,Raffelt:1999tx}. Even with tiny couplings to electrons and protons, they might be still copiously created in the interior of hot and dense stars. Because of their weak interactions they might abandon the star without further scattering, accelerating the consumption of nuclear fuel, and therefore the stellar evolution \\cite{Dicus:1978fp,Frieman:1987ui}. Our present observational data on stellar evolution can strongly constrain this novel luminosity, although the bounds can be relaxed in some concrete models \\cite{Masso:2005ym,Masso:2006gc,Jaeckel:2006id,Jaeckel:2006xm,Brax:2007ak}. Interestingly enough, in this case the Sun itself could be a copious emitter of weakly interacting particles, that could eventually be detected at Earth inside a sensitive detector \\cite{Sikivie:1983ip} (as this is actually the case with neutrinos). Several of these so-called ``helioscopes\" \\cite{vanBibber:1988ge} have been built with the aim of detecting solar axions \\cite{Lazarus:1992ry,Moriyama:1998kd,Zioutas:2004hi} and remarkably, the CAST collaboration has even recently surpassed the sensitivity of the energy loss arguments for the axion coupling to two photons \\cite{Andriamonje:2007ew}. As we will see, helioscopes can also detect hidden photons so the CAST limits can also be used to constrain the solar hidden photon flux. The energy loss argument and helioscope bounds for hidden photons were already studied in \\cite{popov:1991,Popov1999}. However, in this paper the author does not consider either the possibility of a resonant production, which can enhance enormously the hidden photon flux, nor the emission of transversely polarized hidden photons. In this paper we compute the energy loss bounds using the latest solar data \\cite{Bahcall:2004pz} and derive the CAST helioscope bounds. In particular, it is shown that the new CAST results are extremely sensitive to solar hidden photons, providing the strongest constraint of their existence in the mass range $\\muu\\sim 0.01-1$ eV. Moreover, accounting for the resonant production improves the energy loss bounds in \\cite{Popov1999} roughly up to 1 order of magnitude. The paper is organized as follows: In Sec.\\ref{Sproduction} the hidden photon solar emission is derived while the principles of the CAST helioscope detection are reviewed in Sec.\\ref{Shelioscope}. We compute the energy loss and CAST bounds in Sec.\\ref{Sbounds} and finally the conclusions are presented. ", "conclusions": "I have addressed the calculation of the solar emission of a hypothetical hidden sector photon $B^\\mu$ mixing kinetically with the standard model ordinary photon. I have shown that a resonant effect is possible when the dispersion relation of solar plasmons fits the particle-like dispersion relation of the hidden photon. This happens for transverse plasmons if the hidden photon mass $\\muu$ lies in the range $1\\sim295$ eV (the range of the plasma frequency in the solar model used) and for longitudinal plasmons as long as $\\muu\\lesssim295$ eV. The conservative requirement that the hidden photon luminosity should not exceed the solar standard luminosity bounds the amount of kinetic mixing up to $\\mix\\lesssim10^{-14}$ depending on the mass and the polarization. At masses beyond $1$ eV, where the strongest bound is reached, the emission of transversally polarized hidden photons dominates over the emission of longitudinal ones. Below this mass both polarizations contribute in a similar amount. At low masses the bounds are weaker, relaxing proportionally to $\\muu^2$. However, the non observation of a signal in the CAST axion helioscope improves the bounds in this region up to 2 orders of magnitude. Altogether, these are the best limits on the mixing parameter $\\mix$ in the range $3$ meV$<\\muu<40$ keV. A small room for improvement is available for large masses $\\muu\\gg 10$ keV if the CAST detectors were to rise their top energy threshold. It should be interesting to extend this study to other stellar objects like supernovae, white dwarfs, red giants and horizontal branch stars, since these can provide stronger bounds than the Sun, specially at masses where a resonant production is possible." }, "0801/0801.0467_arXiv.txt": { "abstract": "We argue that the typical energy density of a light scalar field should not be less than $H^4$ in the inflationary Universe. This requirement implies that the non-Gaussianity parameter $f_{NL}$ is typically bounded by the tensor-scalar ratio $r$ from above, namely $f_{NL}\\lesssim 518\\cdot r^{1\\over 4}$. If $f_{NL}=10^2$, inflation occurred around the GUT scale. ", "introduction": "Inflation model \\cite{Guth:1980zm,Albrecht:1982wi,Linde:1982zj} provides an elegant mechanism to solve the horizon, flatness and primordial monopole problem due to a quasi-exponential expansion of the universe before the hot big bang. The temperature anisotropies in cosmic microwave background radiation (CMBR) and the large-scale structure of the Universe are seeded by the primordial quantum fluctuations during inflation \\cite{Mukhanov:1990me,Lyth:1998xn}. Since the density perturbation is roughly $10^{-5}$, it is good enough to apply the linear perturbation theory to calculate the quantum fluctuations during inflationary epoch. Within this approach, the Fourier components of fluctuations are uncorrelated and their distribution is Gaussian. That is why the non-Gaussianity from the simplest inflation models is very small ($|f_{NL}|<1$). For useful discussions on non-Gaussianity see \\cite{Salopek:1990jq,Salopek:1990re,Falk:1992sf,Gangui:1993tt,Acquaviva:2002ud,Maldacena:2002vr}, and for a nice review see \\cite{Bartolo:2004if}. The non-Gaussian perturbation is governed by the n-point correlation function of the curvature perturbation \\e \\langle \\Phi({\\bf{k}_1})\\Phi({\\bf{k}_2})\\cdot\\cdot\\cdot\\Phi({\\bf{k}_n})\\rangle=(2\\pi)^3\\delta^3(\\sum_{i=1}^{n}{\\bf{k}_i})F_n(k_1,k_2,...,k_n),\\q where $\\Phi({\\bf{k}})$ is the Fourier mode of Bardeen's curvature perturbation. The leading non-Gaussian features are known as the bispectrum (three-point function) and trispectrum (four-point function), with their sizes conventionally denoted as $f_{NL}$ and $\\tau_{NL}$ respectively. The non-linearity parameter $f_{NL}$ defined in \\cite{Komatsu:2001rj} takes the form \\e \\Phi({\\bf x})=\\Phi_L({\\bf x})+f_{NL}[\\Phi_L^2({\\bf x})-\\langle\\Phi_L^2({\\bf x})\\rangle], \\q where $\\Phi_L$ denotes the linear Gaussian part of the perturbation in real space. The non-Gaussianity parameter $f_{NL}$ characterizes the amplitude of the primordial non-Gaussian perturbations. The value of $\\Phi_L$ is roughly $10^{-5}$. If $f_{NL}=10^2$, the distribution of $\\Phi$ is still consistent with a Gaussian distribution to $0.1\\%$. It is not easy to detect a small non-Gaussianity in the experiments. The shape of non-Gaussianity is also very important. If $F(k_1,k_2,k_3)$ is large for the configurations in which $k_1\\ll k_2,k_3$, it is called local, ``squeezed\" type; if $F(k_1,k_2,k_3)$ is large for the configurations in which $k_1\\sim k_2\\sim k_3$, it is called non-local, ``equilateral\" type. Recently a large primordial local-type non-Gaussianity was reported to be marginally detected from the data of WMAP3 in \\cite{Yadav:2007yy}: \\e 2710$ will rule out most of the existing inflation models. The uncertainty of $f_{NL}^{local}$ from WMAP will shrink to be 42 for 8 years data, and 38 for 12 years data \\cite{Komatsu:kitpc}. The accuracy of PLANCK will be roughly $6$. If the local-type non-Gaussianity is of order $10^2$, it will be detected at high confidence level in the near future. In this paper we investigate the curvaton scenario in detail and find that the inflation scale is roughly at GUT scale in order to get a large local-type non-Gaussianity. Ekpyrotic model can also provide a large local-type non-Gaussianity. However unlike the slightly red-tilted gravitational wave spectrum in the inflation/curvaton model, the gravitational wave spectrum in Ekpyrotic model is strongly blue and then the amplitude is exponentially suppressed on all observable scales \\cite{Boyle:2003km}. Gravitational wave perturbation can be used to distinguish curvaton scenario from Ekpyrotic model. Here we also want to clarify two points in this paper. One is that maybe we miss some order-one coefficients in (\\ref{bsg}) and (\\ref{ggg}), and then the bound in (\\ref{fnlr}) should be modified to be a little looser or more stringent. But the order of magnitude of our result can be trusted. The other is that the bound in (\\ref{fnlr}) depends on the WMAP normalization $P_{\\zeta,{WMAP}}=2.457\\times 10^{-9}$. If we let the normalization of the density perturbation free, we find \\e f_{NL}\\lesssim \\gamma^{-1}\\cdot 518\\cdot r^{1\\over 4}, \\q and eq. (\\ref{fnlp}) becomes\\e f_{NL}\\lesssim \\gamma^{-1}\\cdot 871 \\cdot \\({1\\over \\Delta N_e}{|\\Delta \\phi|\\over M_p}\\)^\\half, \\q where $\\gamma=(P_\\zeta/P_{\\zeta,{WMAP}})^\\half$. Larger the amplitude of the density perturbation, smaller the non-Gaussianity. If we also consider $|\\Delta \\phi|/M_p<1$ for $\\Delta N_e=50$, $f_{NL}\\lesssim 123\\cdot \\gamma^{-1}$. The density perturbation cannot be larger than 10 times of its observed value in our Universe if the non-Gaussianity parameter is not smaller than 10. The equilateral-type non-Gaussianity has not been detected. Many models \\cite{ArkaniHamed:2003uz,Chen:2006nt,ArmendarizPicon:1999rj,Chen:2007gd,Li:2007} were suggested to generate large equilateral-type non-Gaussianity as well. Other mechanisms concerning the large non-Gaussianity are discussed in \\cite{Dvali:2003em,Dvali:2003ar,Matsuda:2007tr,Matsuda:2007ds,Berera:1995ie,Berera:1996nv,Gupta:2002kn}. Anyway, non-Gaussianity will be an important issue for cosmology and string theory. \\vspace{.5cm} \\noindent {\\bf Acknowledgments} We would like to thank M.~Sasaki and P.~J.~Yi for useful discussions." }, "0801/0801.4378_arXiv.txt": { "abstract": "The annihilations of neutralino dark matter (or other dark matter candidate) generate, among other Standard Model states, electrons and positrons. These particles emit synchrotron photons as a result of their interaction with the Galactic Magnetic Field. In this letter, we use the measurements of the WMAP satellite to constrain the intensity of this synchrotron emission and, in turn, the annihilation cross section of the lightest neutralino. We find this constraint to be more stringent than that provided by any other current indirect detection channel. In particular, the neutralino annihilation cross section must be less than $\\approx 3\\times 10^{-26}$cm$^3$/s ($1\\times 10^{25}$cm$^3$/s) for 100 GeV (500 GeV) neutralinos distributed with an NFW halo profile. For the conservative case of an entirely flat dark matter distribution within the inner 8 kiloparsecs of the Milky Way, the constraint is approximately a factor of 30 less stringent. Even in this conservative case, synchrotron measurements strongly constrain, for example, the possibility of wino or higgsino neutralino dark matter produced non-thermally in the early universe. ", "introduction": " ", "conclusions": "" }, "0801/0801.4187_arXiv.txt": { "abstract": "The chemical compositions of seven Carbon-Enhanced Metal-Poor (CEMP) turn-off stars are determined from high-resolution spectroscopy. Five of them are selected from the SDSS/SEGUE sample of metal-poor stars. Another star, also chosen from the SDSS/SEGUE sample, has only a weak upper limit on its carbon abundance obtained from the high-resolution spectrum. The effective temperatures of these objects are all higher than 6000~K, while their metallicities, parametrized by [Fe/H], are all below $-2$; the star with the lowest iron abundance in this study has [Fe/H] = $-3.1.$ Six of our program objects exhibit high abundance ratios of barium ([Ba/H] $> +1$), suggesting large contributions of the products of former AGB companions via mass transfer across binary systems. One star in our study ({\\objf}) exhibits a rapid variation in its radial velocity, which is a strong signature that this star belongs to a close binary. Combining our results with previous studies provides a total of 20 CEMP main-sequence turn-off stars for which the abundances of carbon and at least some neutron-capture elements are determined. Inspection of the [C/H] ratios for this sample of CEMP turn-off stars show that they are generally higher than those of CEMP giants; their dispersion in this ratio is also smaller. We take these results to indicate that the carbon-enhanced material provided from the companion AGB star is preserved at the surface of turn-off stars with no significant dilution, which appears counter to expectations if processes such as thermohaline mixing have operated in unevolved CEMP stars. In contrast to the behavior of [C/H], a large dispersion in the observed [Ba/H] is found for the sample of CEMP turn-off stars, suggesting that the efficiency of the s-process in very metal-poor AGB stars may differ greatly from star to star. Four of the six stars from the SDSS/SEGUE sample exhibit kinematics that are associated with membership in the outer-halo population, a remarkably high fraction. ", "introduction": "\\label{sec:intro} Abundance studies of very metal-poor (VMP; [Fe/H] $ < -2.0$)\\footnote{[A/B] = $\\log(N_{\\rm A}/N_{\\rm B})- \\log(N_{\\rm A}/N_{\\rm B})_{\\odot}$, and $\\log\\epsilon_{\\rm A} = \\log(N_{\\rm A}/N_{\\rm H})+12$ for elements A and B.} stars have been pursued over the past few decades in order to constrain models of nucleosynthesis, stellar evolution, and early chemical enrichment in the Galaxy \\citep[e.g., ][]{beers05}. One important result of these studies is the discovery of Carbon Enhanced Metal-Poor (CEMP) stars, which appear with increasing frequency at lower metallicity \\citep{beers92, beers05, lucatello06, marsteller07}. These stars may be closely related to carbon stars in the Galactic halo, known as CH stars \\citep{keenan42} and subgiant CH stars \\citep{bond74}. Recent chemical abundance studies for CEMP stars have revealed that most (70--80\\%) CEMP stars also exhibit excesses of s-process elements such as Ba (the CEMP-s stars, according to Beers \\& Christlieb 2005), indicating that the origin of the carbon excesses in these stars is likely to be the triple-$\\alpha$ reaction in Asymptotic Giant Branch (AGB) stars \\citep[e.g., ][]{aoki07}. The CEMP stars that are observed at present are likely to have been polluted by the transfer of carbon-enhanced material from a (former) AGB companion across a binary system, while the AGB star itself has now evolved to an unseen white dwarf \\citep[e.g.,][]{lucatello05}. Thus, the abundance patterns of heavy elements in these stars provide useful constraints on models for s-process nucleosynthesis in AGB stars. On the order of 20\\% of CEMP stars exhibit no significant enhancement of their neutron-capture elements (the CEMP-no stars, according to Beers \\& Christlieb 2005), suggesting the existence of other possible origins for their carbon excesses \\citep[e.g., ][]{norris97b, aoki02a}. \\citet{aoki07} have shown that the CEMP-no stars generally occur at very low [Fe/H]; extreme examples of this class of stars include HE~0107--5240 and HE~1327--2326, two hyper metal-poor (HMP) stars with [Fe/H] below $-5.0$ \\citep{christlieb02, frebel05} and having very large carbon excesses ([C/Fe] $\\sim +4$), as well as the recently identified ultra metal-poor (UMP; [Fe/H] = $-4.8$) star HE~0557-4840, with [C/Fe] $= +1.6$ (Norris et al. 2007). Among the CEMP stars for which chemical compositions have been obtained from high-resolution spectroscopy, main-sequence turn-off stars are expected to be of particular importance. In the case of mass transfer in binary systems, the accreted material from the primary AGB star has been mixed at least by the first dredge-up in red giants, while turn-off stars might preserve pure material accreted from the primary at their surfaces. In such cases, one can investigate the efficiency of the production of carbon and neutron-capture elements in AGB stars from abundance measurements of the secondary star. Another interesting view arises from the suggested influence of so-called thermohaline mixing (Charbonnel \\& Zahn 2007; Stancliffe et al. 2007; Denissenkov \\& Pinsonneault 2007), which provides the possibility of mixing the accreted surface material while the observed star is still on the main-sequence or only slightly evolved, prior to first dredge-up. In this scenario, the contrast of the observed surface abundances of CEMP turn-off stars with more evolved CEMP stars also provides valuable clues to the nature of this proposed extensive mixing process. Very large new samples of CEMP stars have recently become available, discovered during the course of the Sloan Digital Sky Survey (SDSS; York et al. 2000; Adelman-McCarthy et al. 2007). Although originally designed as an extragalactic survey, SDSS has also discovered large numbers of VMP stars \\citep{beers06}. Although some of the CEMP stars are the result of directed studies (Margon et al. 2002; Downes et al. 2004), many of them have appeared among the calibration objects used by SDSS to obtain spectrophotometric and telluric corrections for other spectroscopic data. These calibration stars are primarily brighter, metal-poor main-sequence turn-off F- and G-type stars. The ongoing extension to SDSS, SDSS-II (which includes the program SEGUE: Sloan Extension for Galactic Understanding and Exploration), is expected to provide tens of thousands of additional VMP stars, at least several thousand of which are expected to be CEMP stars. This paper reports the first application of abundance measurements obtained with high-resolution spectroscopy for CEMP star candidates found by the SDSS/SEGUE surveys. In \\S 2 we discuss the identification of our sample stars and the observations that were carried out. A description of our analysis techniques and results is provided in \\S 3. In \\S 4 we present a discussion of our findings. The interesting kinematics of the SDSS/SEGUE CEMP turn-off stars are discussed in \\S 5. We conclude with a few remarks in \\S 6. ", "conclusions": "Chemical compositions of seven CEMP turn-off stars are determined. Six stars among them exhibit a large excess of Ba, signature of a contribution by the nucleosynthesis in an AGB star. The distribution of carbon abundances in these stars suggest that the surface of such stars preserves the material transferred from the AGB star that was the erstwhile primary star in a binary system. If this is the case, the relatively wide distribution of Ba abundances ([Ba/H]) indicates a diversity of the efficiency of the s-process in metal-poor AGB stars. Further studies to identify the physical mechanism that produces such diversity are clearly desired. The present study is the first application of high-resolution spectroscopy to candidate CEMP stars from the SDSS and SEGUE sample. Comparisons of our results on stellar parameters and chemical abundances with the estimates from the SDSS spectra confirmed that the selection of metal-poor stars works well in general. The SDSS/SEGUE survey is providing a large sample of candidate metal-poor stars. High-resolution spectroscopy for such stars in the near future will reveal the chemical abundance trends in the lowest metallicity range, as well as be useful for exploring the possible dependence of their chemical properties on their derived kinematics." }, "0801/0801.3591_arXiv.txt": { "abstract": "{Deep optical CCD images of the supernova remnant G 15.1$-$1.6 were obtained and filamentary and diffuse emission has been discovered. The images, taken in the emission lines of \\hnii, \\sulfur~and \\oiii, reveal filamentary and diffuse structures all around the remnant. The radio emission at 4850 MHz in the same area is found to be well correlated with the brightest optical filaments. The IRAS 60$\\mu$m emission may also be correlated with the optical emission but to a lesser extent. The flux calibrated images suggest that the optical emission originates from shock-heated gas (\\sulfur/\\ha\\ $>$ 0.4), while there is a possible \\HII\\ region (\\sulfur/\\ha\\ $\\sim$ 0.3) contaminating the supernova remnant's emission to the east. Furthermore, deep long--slit spectra were taken at two bright filaments and also show that the emission originates from shock heated gas. An \\oiii\\ filamentary structure has also been detected further to the west but it lies outside the remnant's boundaries and possibly is not associated to it. The \\oiii\\ flux suggests shock velocities into the interstellar $``$clouds'' $\\sim$100 \\vel, while the \\siirat\\ ratio indicates electron densities up to $\\sim$250 cm$^{-3}$. Finally, the \\ha\\ emission has been measured to be between 2 to 7 $\\times$ \\fluxa, while the lower limit to the distance is estimated at 2.2 kpc. ", "introduction": "Supernova remants (SNRs) play an important role to understand the SN mechanism, the interstellar medium (ISM) and their interaction. Most of the SNRs have been detected in radio from their non--thermal synchrotron emission. Observations of SNRs in X--rays allow us to directly probe the hot gas inside the primary shock wave, while optical observations offer an important tool for the study of the interaction of the shock wave with dense material found in the ISM. New searches in optical waveband continue to identify Galactic SNRs (e.g. Boumis et al. \\cite{bou02}, \\cite{bou05}; Mavromatakis et al. \\cite{mav02}, \\cite{mav05}) while in the last decade, observations in X--rays have also detected new Galactic SNRs (e.g. Seward et al. \\cite{sew95}; see also Green \\cite{gre06} for a complete catalogue). G 15.1--1.6 is not a well known SNR, and was first detected by Reich et al. (\\cite{rei88}) in the Effelsberg 2.7 GHz survey and the radio image was published by Reich et al. (\\cite{rei90}). It is classified as a shell--type SNR, with a spectral index of $\\sim$0.8. Its angular size is 30\\arcmin $\\times$ 24\\arcmin and using the brightness--to--diameter ($\\Sigma$--D) relationship the distance of the remnant was calculated at 5.7 kpc (Green \\cite{gre06}). Radio surveys of the surround area do not reveal any pulsar to by associated with G15.1--1.6 while it has not been detected optically in the past. In this paper we report the optical detection of G 15.1--1.6. We present images of the remnant in the \\hnii, [S~{\\sc ii}] and [O~{\\sc iii}] emission lines. Deep long slit spectra were also acquired in a number of selected areas. In Sect. 2 we present information about the observations and data reduction, while the results of the imaging and spectroscopic observations are given in Sect. 3. In Sect. 4 we discuss the optical properties of this SNR, while in Sect. 5 we summarize the results of this work. ", "conclusions": "The faint supernova remnant G 15.1$-$1.6 was observed for the first time in major optical emission lines. The images show filamentary and diffuse emission structures. The bright filaments are very well correlated with the remnant's radio emission at 4850 MHz suggesting their association, while correlation evidence also shown with the IRAS 60$\\mu$m map. The flux calibrated images and the long--slit spectra indicate that the emission arises from shock heated gas. Finally, an upper limit for the electron density of 250 \\dens and a lower limit for the distance of 2.2 kpc are calculated." }, "0801/0801.1594_arXiv.txt": { "abstract": "{ We present a compilation of measurements of the stellar mass density as a function of redshift. Using this stellar mass history we obtain a star formation history and compare it to the instantaneous star formation history. For $z<0.7$ there is good agreement between the two star formation histories. At higher redshifts the instantaneous indicators suggest star formation rates larger than that implied by the evolution of the stellar mass density. This discrepancy peaks at $z=3$ where instantaneous indicators suggest a star formation rate around $0.6$ dex higher than those of the best fit to the stellar mass history. We discuss a variety of explanations for this inconsistency, such as inaccurate dust extinction corrections, incorrect measurements of stellar masses and a possible evolution of the stellar initial mass function. } ", "introduction": "Much contemporary research in extragalactic astronomy has revolved around the determination of the instantaneous cosmic star formation history (SFH, Madau et al. 1996; Lilly et al. 1996). However, measuring this quantity from observations requires a number of assumptions, with the form of the dust obscuration corrections and stellar initial mass function (IMF, see Kroupa 2007a for a recent overview) being among the most important. Integration of the instantaneous star formation history over redshift, making appropriate corrections for stellar evolution processes, yields the current stellar mass density. This quantity can be independently measured, typically using extensive galaxy surveys such as the 2dFGRS or SDSS, often combined with near infrared (NIR) measurements. Numerous studies have attempted comparisons of these quantities. Madau, Pozzetti \\& Dickinson (1998), Cole et. al (2001), Fontana et al. (2004) and Arnouts et al. (2007) all found good agreement between the SFH with a low dust content and measured values of the stellar mass density. On the other hand there have been a number of studies (Eke et al. 2005; Hopkins \\& Beacom 2006, hereafter HB06) which claim that the instantaneous SFH overpredicts the low redshift stellar mass density. We attempt to investigate this possible discrepancy using a compilation of the most up to date measurements of the stellar mass density history (SMH, the $\\Omega_{*}$-redshift relation). This relation is intricately connected to the instantaneous star formation history, but in this context it has some important advantages. The principal advantage is that estimates of stellar mass typically probe a range of the stellar mass function that is somewhat more representative of the stellar mass whereas instantaneous indicators probe only the most massive stars. Furthermore instantaneous measurements can be subject to a greater uncertainty introduced by the effects of dust obscuration. In \\S 2 we present a compilation of both low and high redshift measurements of the stellar mass density. Using these values, in \\S 3 we derive a best fitting star formation history. We then compare this estimate in \\S 4 to other estimates of the star formation history and highlight any discrepancies. Finally, in \\S 5 we present a discussion of our results and end in \\S 6 with a summary. Throughout this work, we assume a flat $\\Lambda$ CDM cosmology with $\\Omega_{\\Lambda}=0.7$, $\\Omega_{\\rm matter}=0.3$ and $H_{0} = 70\\,\\, {\\rm kms}^{-1}\\,Mpc^{-1}.$ ", "conclusions": "In this work a compilation of stellar mass density measurements over the range $0$ 3000 \\kms), so as to understand their origin. Normalized, composite spectra were derived, for absorption line measurements, for the full sample and for several sub-samples, chosen on the basis of the line strengths and other absorber and QSO properties. Composite absorption lines differ in small but measurable ways from those in the composite spectra of intervening absorption line systems, especially in the relative strengths of {Si~IV}, {C~IV} and {Mg~II}. From the analysis of the composite spectra, as well as from the comparison of measured equivalent widths in individual spectra, we conclude that the associated {Mg~II} absorbers have higher apparent ionization, measured by the strength of the {C~IV} absorption lines compared to the {Mg~II} absorption lines, than the intervening absorbers. The ionization so measured appears to be related to apparent ejection velocity, being lower as the apparent ejection velocity is more and more positive. Average extinction curves were obtained for the sub-samples by comparing their geometric mean QSO spectra with those of matching (in \\zem $~$and \\imag) samples of QSOs without absorption lines in their spectra. There is clear evidence for dust-like attenuation in these systems, though the 2175 {\\AA} absorption feature is not present: the extinction is similar to that found in the Small Magellanic Cloud. The extinction is almost twice that observed in the similarly selected sample of intervening systems. We reconfirm with our technique that QSOs with non-zero FIRST (Faint Images of the Radio Sky at Twenty-cm) radio flux are intrinsically redder than the QSOs with no detection in the FIRST survey. The incidence of associated {Mg~II} systems in QSOs with non-zero FIRST radio flux is 1.7 times that in the QSOs with no detection in the FIRST survey. The associated absorbers in radio-detected QSOs which comprise about 12\\% of our sample, cause three times more reddening than the associated absorbers in radio-undetected QSOs. The origin of this excess reddening in the absorbers is indicated by the correlation of the reddening with the strength of Mg II absorption. This excess reddening possibly suggests an intrinsic nature for the associated absorbers in radio-detected QSOs. ", "introduction": "Many of the first detected, narrow-line QSO absorption systems had redshifts close to those of their QSOs (e.g. Stockton \\& Lynds 1966; Burbidge, Lynds \\& Burbidge 1966). Such systems, having a relative velocity with respect to the QSO, in units of the speed of light\\footnote{$\\bet={{(1+\\zem)^2-(1+\\zab)^2}\\over{(1+\\zem)^2+(1+\\zab)^2}}$}, $\\beta$, smaller than 0.02, are now termed associated systems. Study of the associated systems is important from the point of view of understanding the energetics and kinematics near the central black hole and also for understanding the ionization structure, dust content and abundances in material directly exposed to the radiation from the QSOs, and in some cases, possibly ejected from, the QSOs or the accretion disks. Associated systems have been suggested to arise (a) in the outer parts of QSO host galaxies (e.g. Heckman et al. 1991; Chelouche et al. 2007) which may have gas properties similar to those in the outer parts of inactive galaxies (Steidel et al. 1997, Fukugita \\& Peebles 2006); (b) by material within 30 kpc of the AGN, accelerated by starburst shocks from the inner galaxy (e.g. Heckman et al. 1990, 1996; D'Odorico et al. 2004; Fu and Stockton 2007a); (c) in the core of the AGN, within 10 pc of the black hole (Hamann et al. 1997a; Hamann et al. 1997b; Barlow \\& Sargent 1997). In case (a), these are possibly ``halo\" clouds as in normal galaxies, possibly lit up by the QSO to produce extended regions of Lyman alpha emission, but not thought to be moving at the high velocities necessary to explain the dispersion of associated system velocities with respect to the QSO itself. However, using spectra of a set of angularly close QSO pairs, Bowen et al. (2006) confirmed that gas in the outer parts of some QSO host galaxies is detectable as absorption in the spectra of background QSOs, but the spectra of the foreground QSOs did not reveal associated absorption systems: evidently, the appearance of the absorption is dependent on the angle of the line of sight to the spin axis of the accretion disk. The low velocities expected from such gas led to the postulate that clusters of galaxies near the QSO showing associated absorption were responsible for the absorption and the dispersion of cloud velocities. However, QSOs are not only associated with clusters of galaxies: they appear in a wide range of galaxy types and masses (Jahnke et al. 2004) and, while found in slightly higher density environments (Serber et al. 2006), do not require cluster type densities on large scales (Wake et al. 2004). In case (b), the gas is typically found to have densities of a few hundred particles cm$^{-3}$, high enough to produce excited fine structure lines of {Si~II} and {C~II}. There are suggestions in the above references that this is material ejected by a galactic wind, possibly a superwind from a starburst (Heckman et al. 1990), then lit up by the QSO. (These are often called extended emission line regions, EELR). Cooling flows (Crawford \\& Fabian 1989) no longer seem to be considered in most cases (Fu \\& Stockton 2007b). In case (c), absorption is inferred to be very close to the black hole because of variability in the absorption lines, and/or because of the presence of clouds that do not cover the source. Theoretical investigations of case (c) (Arav et al. 1994; Konigl \\& Kartje 1994; Murray et al. 1995; Krolik \\& Kriss 2001; Proga et al. 2000; Everett 2005; Chelouche \\& Netzger 2005 (for a Seyfert galaxy with a much lower radiation field)), have reinforced the plausibility of thermal or hydromagnetic, radiation assisted flows, both parallel to the AGN jet axis, and perpendicular to it, along the accretion disk. See, for example, Figure 13 of Konigl \\& Kartje (1994), Figure 1 of Richards et al. (1999), Figure 1 of Murray et al. (1995) and Figure 1 of Everett (2005). Early studies concentrated on systems with C IV doublets. Weymann et al. (1979) and Foltz et al. (1986) found a statistical excess of such systems as compared to what was expected if these were randomly distributed in space. Other studies (Young et al. 1982; Sargent et al. 1988) could not confirm these observations. Later it was shown that {Mg~II} systems (Aldcroft et al. 1994) and C IV systems (Anderson et al. 1987; Foltz et al. 1988) with \\bet $<$ 0.0167 are preferentially found in steep-spectrum radio sources, often thought of as sources dominated by emission from lobes rather than the core of the source. Ganguly et al. (2001) showed that high ionization systems (having lines of C IV, N V and O VI) with \\bet $<$ 0.0167 with $z_{abs}<1.0$ are not present in radio-loud QSOs that have compact radio morphologies, flat radio spectra (core dominated sources) and C IV lines with mediocre FWHM ($\\le 6000$ \\kms). Baker et al. (2002) from a study of a near-complete sample of low-frequency selected, radio-loud QSOs corroborated trends for C IV associated absorption to be found preferentially in steep-spectrum and lobe-dominated QSOs, suggesting that the absorption is the result of post-starburst activity and that the C IV lines weaken as the radio source grows, clearing out the gas and dust. Vestergaard (2003, hereafter V03) studied a sample of high ionization associated systems with $1.5$ 0.01) selected with similar criteria. In particular, we are looking for a spectroscopic signature to distinguish associated from intervening systems. We make use of the composite spectra of the sample (and various sub-samples thereof), following the method recently advocated by York et al. (2006, hereafter Y06). In Section 2, we describe the criteria used for sample selection, various sub-samples generated from the main sample and the method of generating composite spectra. In section 3, we present our results which include (i) a comparison between the properties of the samples of the intervening and associated systems, using several statistical tests; (ii) a discussion of the line strengths in the composite spectra of associated systems compared to those for the intervening sample and of the state of ionization in these systems; (iii) the measured extinction for various sub-samples; (iv) a detailed analysis of the dependence of extinction and other properties on the radio properties of the QSOs; (v) discussion of the abundances in associated sample; and (vi) possible scenarios for the origin and location of the absorbers. A few systematic effects that may be hidden in our data are noted in section 4. Conclusions are presented in section 5. ", "conclusions": "We have studied a sample of 415 intermediate redshift, {Mg~II} associated systems in the spectra of SDSS QSOs. Our main conclusions are as follows. \\begin{enumerate} \\item There is definite evidence of dust extinction in the associated systems. The average extinction is two times that found in a sample of similarly selected, intervening systems. This larger extinction could be attributed to higher average N$_{\\rm H\\;I}$ +N$_{\\rm H\\;II}$ in associated compared to intervening systems, or to differences in the dust to gas ratio. \\item There is no evidence for the 2175 {\\AA} bump in the extinction curves of the associated absorbers. \\item The extinction curve for these absorbers is similar to that of the SMC, which has also been found to be the case for intervening absorbers. \\item Associated absorbers are in a higher state of ionization compared to the intervening absorbers. \\item In the relative velocity range 0.01$>$\\bet$>$-0.004 studied, the ionization conditions and the total extinction in the associated systems is a function of their apparent relative velocity with respect to the QSO, systems with lower relative velocity being more ionized and more highly reddened. \\item There is no obvious evidence for higher abundances in the associated systems than those in the intervening systems. \\item About a third of the absorption systems do have redshifts higher than the emission redshift of QSOs and thus appear to be infalling. The direct inference would be that lower \\bet $~$systems are closer to the ionization sources. No clues are found that distinguish the material as ambient material (shocked?) in the QSO host galaxies or as returning gas from QSO jet or accretion disk outflows. \\item The QSOs with non-zero radio flux in the FIRST survey are intrinsically redder than QSOs with null detection in the FIRST survey by relative \\ebv $\\sim$0.04, on average, for an SMC extinction curve, even when there are no absorption lines present. \\item Associated systems in QSOs with non-zero radio flux in the FIRST survey have three to four times as much dust as those in QSOs with null detection in the FIRST survey. This excess reddening is correlated to the strength of Mg II absorption lines and appears to originate in the absorbers themselves. These absorbers are thus, significantly different from the intervening systems and may possibly be intrinsic to the QSOs. \\item No clear discriminant between associated and intervening systems has been found that would definitely work on a case by case basis, despite the clear but subtle differences in ionization and reddening noted above. ``Intrinsic\" but high ejection velocity systems may be hard to discern among the intervening systems, except through non-detection of a galaxy at the absorber redshift. \\end{enumerate} \\appendix" }, "0801/0801.2903_arXiv.txt": { "abstract": "We study the spectral energy distribution of gamma rays and neutrinos in the precessing microquasar SS433 as a result of $pp$ interactions within its dark jets. Gamma-ray absorption due to interactions with matter of the extended disk and of the star is found to be important, as well as absorption caused by the UV and mid-IR radiation from the equatorial envelopment. We analyze the range of precessional phases for which this attenuation is at a minimum and the chances for detection of a gamma-ray signal are enhanced. The power of relativistic protons in the jets, a free parameter of the model, is constrained by HEGRA data. This imposes limits on the gamma-ray fluxes to be detected with instruments such as GLAST, VERITAS and MAGIC II. A future detection of high energy neutrinos with cubic kilometer telescopes such as IceCube would also yield important information about acceleration mechanisms that may take place in the dark jets. Overall, the determination of the ratio of gamma-ray to neutrino flux will result in a key observational tool to clarify the physics of heavy jets. ", "introduction": "The famous and enigmatic microquasar SS433 has been matter of investigation for more than two decades. Consisting of a donor star feeding mass to a black hole, it presents two oppositely directed, precessing jets with hadronic content\\footnote{Iron lines with a shift corresponding to a velocity of $v\\sim 0.26 c$ have been detected, for instance, by Migliari et al. (2002).}. We refer to the relativistic collimated outflows as `dark' jets \\citep{gallo} since the very high kinetic luminosity $L_{\\rm k}\\sim 10^{39}$ erg s$^{-1}$ \\citep{dubner} appears as the dominant power output of the ejected material, having imprinted a deformation on the supernova remnant W50. % Most of the radiative output of the system is observed in the UV and optical bands, whereas the X-ray emission detected is $\\sim 1000$ lower than the kinetic energy of the jets, probably due to a screening effect with an equatorial outflow \\citep{fabrikaXrays,marshall}. The gamma-ray emission above $0.8$ TeV has been constrained by HEGRA to be $\\Phi_\\gamma<8.93 \\times 10^{-13} {\\rm cm}^{-2}{\\rm s}^{-1}$ \\citep{HEGRA} whereas the neutrino flux upper limit according to AMANDA-II data is $\\Phi_\\nu<0.21 \\times 10^{-8} {\\rm cm}^{-2}{\\rm s}^{-1}$ \\citep{Halzen06}. In previous hadronic models for high energy emission from microquasars, relativistic protons in the jets interact with target protons from the stellar wind of the companion star \\citep{rom03,hugo02,Or07}. Since in the case of SS433 there is no evidence of such a strong stellar wind, in this work we investigate the possible production of gamma rays and neutrinos resulting from $pp$ interactions between relativistic and cold protons within the jets themselves. \\section[]{Preliminaries} The binary SS433, distant $5.5$ kpc from the Earth, displays two mildly relativistic jets ($v_{\\rm b}\\approx 0.26$c) that are oppositely directed and precess in cones of half opening angles of $\\theta\\approx 21^\\circ$. The line of sight makes an angle $i=78^\\circ$ with the normal to the orbital plane and a time-dependent angle $i_{\\rm j}(t)$ with the approaching jet (see Fig. {\\ref{Figrender}}). Assuming that $\\psi(t)$ is the precessional phase of the approaching jet, we shall follow the convention that when $\\psi=0$ the mentioned jet points closer to the Earth. Then, when $\\psi=0.5$, it has performed half of the precession cycle and it makes its largest angle with the line of sight. The mass loss rate in the jets is $\\dot{m}_{\\rm j}= 5\\times 10^{-7}{ M}_\\odot {\\rm yr}^{-1}$, the period of precession is $162$ d and the orbital period is $13.1$ d \\citep{Fabrika100}. The donor star and the compact object are thought to be embedded in a thick expanding disk which is fed by a wind from the supercritical accretion disk around the black hole \\citep{Zwitter}. This equatorial envelope is perpendicular to the jets and according to \\citet{Fabrika100} we assume that it has a half opening angle $\\alpha_{\\rm w} \\approx 30^\\circ $, a mass loss rate $\\dot{M}_{\\rm w}\\approx 10^{-4}{ M}_\\odot{\\rm yr}^{-1} $ and a terminal velocity $v_{\\rm w}\\sim 1500 {\\rm \\ km \\ s}^{-1}$. Also, this extended disk has been recognized as the origin of both the UV and mid-IR emission \\citep{Gies02a,Fuchs05} which can cause significant absorption of gamma-rays as discussed in \\citet{last}. The spectral identification of the companion star has been difficult due to the presence of the extended disk, since the star is often partially or totally obscured by it. After convenient observations at specific configurations of precessional and orbital phases it has became quite clear that the star is an A-supergiant \\citep{Hillwig,Barnes,Chere}. We assume the masses of the components as derived from INTEGRAL observations \\citep{Chere}, $M_{\\rm bh}= 9 { M}_\\odot$ and $M_\\star= 30 { M}_\\odot$ for the black hole and star respectively. This corresponds to an orbital separation $a \\simeq 79 \\ { R}_\\odot$ for a zero-eccentricity orbit as it is the case for SS433. Since the star is believed to fill its Roche lobe, the implied radius according to \\citet{Eggleton} is $R_L\\simeq 38 { R}_\\odot$. \\begin{figure} \\includegraphics[trim = 0mm 0mm 0mm 0mm, clip, width=8cm,angle=0]{MN-08-0074-MJ-Fig1.eps} \\caption{Schematic view of the SS433. The \\textit{approaching} jet is most of the time closest to our line of sight and the \\textit{receding} one is oppositely directed.} \\label{Figrender} \\end{figure} \\subsection{Outline of the jet model} \\textbf{} We assume that a magneto-hydrodynamic mechanism for jet ejection operates in SS433, that is, ejection is realized through the conversion of magnetic energy into matter kinetic energy. The magnetic energy density is supposed to be in equipartition with the kinetic energy density of the ejected particles, so that the corresponding magnetic field along the jet is given by \\be B(z_{\\rm j})= \\sqrt{8\\pi e_{\\rm j}}, \\ee where the kinetic energy density is \\be e_{\\rm j}=\\frac{\\dot{m}_{\\rm j}E_{\\rm k}}{m_{\\rm p}v_{\\rm b}\\pi R_{\\rm j}^2(z_{\\rm j})}. \\ee Here $E_{\\rm k}$ is the classical kinetic energy of a jet proton with velocity $v_{\\rm b}$ and $R_{\\rm j}(z)$ is the jet radius at the height $z_{\\rm j}$ along the jet axis. The jets are modeled as cones with a half opening angle $\\xi_{\\rm j}\\approx 0.6^\\circ$ \\citep{marshall}. Assuming an initial jet radius $R_0=R_{\\rm j}(z_0)\\approx 5 R_{\\rm Sch}$, where $R_{\\rm Sch}=2GM_{\\rm bh}/{c^2}$, we find the injection point as $z_0= R_0/\\tan{\\xi_{\\rm j}}\\simeq 1.3 \\times 10^{9}$cm. Since the jets are heavy as compared to other similar objects, it is reasonable to admit that they are cold matter dominated. In this case, we assume that a small fraction of relativistic or hot particles are confined by the cold plasma. According to \\citet{broadband} the pressure of cold particles is greater than that of the relativistic ones if the ratio of cold to hot particles is less than 1/1000, and this condition will be greatly satisfied provided that the luminosity carried by relativistic particles is required to be smaller than the total kinetic luminosity of the jet. Particle acceleration is supposed to take place via diffusive acceleration by internal shocks converting bulk kinetic energy into random kinetic energy. According to the standard model for non-relativistic shock acceleration (e.g. Blandford \\& Eichler 1987 and references therein) we expect that the relativistic proton spectrum is given by a power-law, $N'_p({E'}_p)= K_0 {E'}_p^{-\\alpha}$ at $z_{\\rm j}=z_0$, where the spectral index is the standard value for first order diffusive shock acceleration, $\\alpha=2$. The flux of these protons hence evolves with $z_{\\rm j}$ as $$J'_p({E'}_p)= \\frac{cK_0}{4\\pi} (z_0/z_{\\rm j})^2 {E'}_p^{-\\alpha}$$ in the jet frame, which transformed to the observer frame \\citep{Purmohammad} reads \\begin{multline} J_p(t,E_p,z_{\\rm j})=\\frac{c K_0}{4 \\pi} \\left(\\frac{z_0}{z_{\\rm j}}\\right)^ 2 \\times \\label{Jplab} \\\\ \\frac{\\Gamma^{-\\alpha+1} \\left(E_p-\\beta_{\\rm b} \\sqrt{E_p^2-m_p^2c^4} \\cos i_{\\rm j} \\right)^{-\\alpha}}{\\sqrt{\\sin ^2 i_{\\rm j} + \\Gamma^2 \\left( \\cos i_{\\rm j}(t) - \\frac{\\beta_{\\rm b} E_p}{\\sqrt{E_p^2-m_p^2 c^4}}\\right)^2}} \\\\ \\equiv \\left(\\frac{z_0}{z_{\\rm j}}\\right)^2 \\tilde{J}_p(E_p,t), \\end{multline} where $i_{\\rm j}(t)$ is the angle between the jet axis and the line of sight, $\\beta_{\\rm b}= 0.26$, and $\\Gamma=\\left[ 1-\\beta_{\\rm b}^2 \\right]^{-1/2}$ is the jet Lorentz factor. The normalization constant $K_0$ is obtained by specifying the fraction of power carried by the relativistic protons, $q_{\\rm rel}$, \\be \\pi R_0^2\\int_{E_p'^{\\rm(min)}}^{E_p'^{\\rm(max)}} J'_p(E'_p) E'_p dE'_p= q_{\\rm rel}L_k, \\ee so that \\be K_0 = \\frac{4 q_{\\rm rel} L_k}{c R_0^2\\ln \\left(\\frac{{E'}_{p}^{\\rm(max)}}{{E'}_p^{\\rm(min)}}\\right)}, \\label{K0norm} \\ee where we take ${E'}_p^{\\rm(min)}\\approx 1$ GeV and the maximum proton energy ${E'}_p^{\\rm(max)}$ will be determined in the next section. We shall adopt, for the illustrative predictions of neutrino and gamma-ray fluxes, a tentative value $q_{\\rm rel}= 10^{-4}$, but a full discussion of the possible range for this parameter will be presented in Sect. \\ref{neutrinos}. ", "conclusions": "We have studied the high-energy emission originated in the dark jets of the microquasar SS433. A small fraction of its particle contents are relativistic protons that collide with the cold ions within the jets, producing gamma rays and neutrinos after pion decay processes. We found that up to distances $\\sim 10^{12}{\\rm cm}$ from the black hole, protons with energies below $\\sim 3\\times 10^{6}$GeV will cool dominantly via $pp$ interactions. The ratio of the power carried by relativistic protons to the total kinetic power of the jet, $q_{\\rm rel}$, was kept as a free parameter of the model (for illustrative purposes we have used $q_{\\rm rel}= 10^{-4}$ in the figures). We have calculated the spectrum of gamma rays and high-energy neutrinos based on the formulae given by \\citet{Kelner06}. We have considered the cooling of the charged pions and muons produced, and we have found that the high-energy neutrino emission is attenuated by synchrotron losses. Adding the contribution from both jets, we have obtained the total gamma-ray spectral intensity of SS433. The gamma radiation will be largely absorbed while leaving the inner regions the system by means of several processes. This will mainly occur through interaction with matter of the star and extended disk, leading to significant photopair production. UV and mid-IR photons originated in the extended disk are also expected to cause important absorption via $\\gamma\\gamma$ annihilations. The total optical depth is found to depend on the precessional phase in such a way that when the approaching jet is pointing away from the Earth, at $\\psi\\sim 0.5$, the extended disk blocks the emitting region and the absorption is strongest. In particular, in the range of precessional phases between $\\psi \\gtrsim 0.91$ and $\\psi \\lesssim 0.09$, the gamma rays originated at the base of the jets will travel to the Earth without having to pass through the equatorial disk. With the conservative assumption that this outflowing disk presents a large half opening angle $\\alpha_{\\rm w}=30^\\circ$, the mentioned range of favorable precessional phases gives a total of $\\sim 29$ d for optimal detectability. Since, according to \\citet{Gies02b}, $\\psi=0$ occurred on 2002 June 5, it follows that the next upcoming opportunities to achieve detection will be centered around the following dates every $162$ d: 2008 August 20, 2009 January 29, 2009 July 10, etc. As mentioned above, the exact duration along which the favorable conditions may hold, depends on the half opening angle of the extended disk, which might be smaller than what was assumed. In that case, the observational window would be broader The observations by HEGRA imply a maximum value for the free parameter $q_{\\rm rel}$ in $\\sim 3 \\times 10^{-4}$. Given the expected sensitivity of the next km$^3$ neutrino telescopes generation, it will be possible to test our model % down to values $\\sim 5\\times 10^{-5}$ in three years of operation. An extended range of this parameter will be probed by gamma-ray observations with GLAST and Cherenkov telescopes, especially if performed on the favorable dates. We conclude that its dark jets can be the possible site for both gamma-ray and neutrino production in SS433. Since most of the high-energy flux is generated in the inner jets, gamma-ray absorption will make detection with Cherenkov telescopes like VERITAS and MAGIC-II difficult but not impossible if attempted at the favorable dates when the approaching jet is closest to the line of sight. Actually, there are much better prospects for gamma-ray observation with GLAST and neutrino detection with IceCube also seems promising. % The determination of the gamma-ray to neutrino flux ratio would allow to estimate $q_{\\rm rel}$ unambiguously, yielding crucial information about acceleration mechanisms taking place in the jets." }, "0801/0801.2768_arXiv.txt": { "abstract": "Mass segregation stands as one of the most robust features of the dynamical evolution of self-gravitating star clusters. In this paper we formulate parametrised models of mass segregated star clusters in virial equilibrium. To this purpose we introduce mean inter-particle potentials for statistically described unsegregated systems and suggest a single-parameter generalisation of its form which gives a mass-segregated state. We describe an algorithm for construction of appropriate star cluster models. Their stability over several crossing-times is verified by following the evolution by means of direct $N$-body integration. ", "introduction": "Observations show quite often an increased concentration of massive stars towards the centres of young star clusters (e.g. ONC -- Hillenbrand \\& Hartmann~1998; NGC 2157 -- Fischer et al.~1998; NGC 3603 -- Stolte et al.~2006). This tendency, known as mass segregation, can be of different origin: Initial mass segregation is sometimes considered (e.g. Murray \\& Lin~1996, Bonnell \\& Bate~2006) as a consequence of the formation of massive stars preferably in the densest regions (i.e. the cores) of the parent gas clouds. On the other hand, the process of mass segregation is also known to be one of the most robust features of the two-body relaxation driven evolution of self-gravitating star clusters (Chandrasekhar~1942, Spitzer~1969). Several approaches were developed to setup a star cluster in the state of mass segregation. Gunn \\& Griffin~(1978), Capuzzo Dolcetta et. al~(2005) and others based their setup on multi-component King models (King~1965, Da~Costa \\& Freeman~1976) with stars separated into several mass classes which interact with each other via smoothed potentials. This approach relies on solving of non-linear set of Poisson equations, which is possible for limitted number of components. A multimass models of star cluster with exact energy equipartition in the core, which also leads to mass segregation, was introduced by Miocchi~(2006). Another approach used e.g. by McMillan \\& Vesperini~(2007) relies on segregation produced by N-body integration of initially unsegregated systems towards the segregated state, i.e. it is equivalent to a simple redefinition of time $t=0$. In this paper we describe a new class of models of star clusters with continuous stellar mass distributions and a parametrised degree of mass segregation. The models are motivated by a study of the process of mass segregation during dynamical evolution of a self-gravitating cluster, which is briefly described in the following Section. We show that mass segregation strongly manifests itself in the energy space. In Section~\\ref{sec:model} we introduce convenient characteristics of a statistically described ensemble and derive their form for the unsegregated state. We further introduce in a heuristic manner an alternative, single-parameter form of these quantities that gives constraints on the distribution function of a mass segregated system. Afterwards, we describe an algorithm for construction of the corresponding star cluster. In Section~\\ref{sec:tests} we demonstrate the stability of the models by means of $N$-body integrations. Finally, Section~\\ref{sec:conclusions} contains our conclusions. ", "conclusions": "\\label{sec:conclusions} We have introduced a way of parametrisation of self-gravitating systems in terms of mean inter-particle potentials. We have demonstrated that this approach can be used for construction of quasi-stationary models of mass segregated star clusters. For the sake of simplicity, we have performed tests with simple power-law mass function. Nevertheless, the approach does not depend on a particular form of the mass function and the standard IMF (Kroupa~2001, 2007) can be used as an input. Notice also that even for a cluster of equal mass stars the algorithm will lead to a system with a desired level of `energy segregation' which is in general the process that drives the star clusters towards core collapse. Finally, let us remark that the index of mass segregation is likely to be related to the entropy. For $S=0$ the system is highly symmetric in terms of $\\mmean{U^{ij}}$. On the other hand, in the limit of $S=1$ it is required that all binding energy is carried by the two most massive particles, which is usually considered a state of maximal entropy of the self-gravitating system. We suggest that the statistical approach based on characterisation of the system by mean values of suitable physical quantities related to subsets of stars deserves further investigation, providing us, hopefully, with a deeper understanding of the thermodynamics of star clusters." }, "0801/0801.2898_arXiv.txt": { "abstract": "Recent evidence of a young progenitor population for many Type-Ia SNe (SNe-Ia) raises the possibility that evolved intermediate-mass progenitor stars may be detected in pre-explosion images. NGC~1316, a radio galaxy in the Fornax cluster, is a prolific producer of SNe-Ia, with four detected since 1980. We analyze Hubble Space Telescope (HST) pre-explosion images of the sites of two of the SNe-Ia that exploded in this galaxy, SN2006dd (a normal Type-Ia) and SN2006mr (likely a subluminous, 1991bg-like, SN-Ia). Astrometric positions are obtained from optical and near-IR ground-based images of the events. We find no candidate point sources at either location, and set upper limits on the flux in $B$, $V$, and $I$ from any such progenitors. We also estimate the amount of extinction that could be present, based on analysis of the surface-brightness inhomogeneities in the HST images themselves. At the distance of NGC~1316, the limits correspond to absolute magnitudes of $\\sim -5.5$, $-5.4$, and $-6.0$ mag in $M_B$, $M_V$, and $M_I$, respectively. Comparison to stellar evolution models argues against the presence at the SN sites, 3 years prior to the explosion, of normal stars with initial masses $\\ga 6M_{\\odot}$ at the tip of their asymptotic-giant branch (AGB) evolution, young post-AGB stars that had initial masses $\\ga 4M_{\\odot}$, and post-red-giant stars of initial masses $\\ga 9 M_{\\odot}$. ", "introduction": "Although there is general consensus that Type-Ia supernova (SN-Ia) explosions are the result of a thermonuclear runaway in a degenerate, approximately Chandrasekhar-mass ($M_{\\rm ch}$), carbon-oxygen stellar core (e.g., Hillebrandt \\& Niemeyer 2000; Badenes et al. 2006; Mazzali et al. 2007), the progenitor systems of SNe-Ia are still unknown. The two main models leading to a SN-Ia explosion are the single-degenerate (SD) model, in which a white dwarf (WD) accretes matter from a close companion until it approaches the Chandrasekhar limit (Whelan \\& Iben 1973; Nomoto 1982), and the double-degenerate (DD) model, involving the merger of two WDs (Iben \\& Tutukov 1984; Webbink 1984). However, both scenarios face theoretical and observational challenges. The outcome of a DD merger may be accretion-induced core collapse, rather than a SN-Ia (e.g., Nomoto \\& Iben 1985; Saio \\& Nomoto 2004, Guerrero et al. 2004), although there are opposing views that invoke rotation of the stellar surface to prevent this outcome (e.g., Piersanti et al. 2003). An observational search for DD progenitor systems (Napiwotzki et al. 2004; Nelemans et al. 2005) has turned up, among $\\sim 1000$ WDs surveyed, few or no potential DD systems with a total mass exceeding $M_{\\rm ch}$ that will merge within a Hubble time. The SD scenario, in turn, has been criticized (e.g., Cassisi et al. 1998; Piersanti et al. 1999, 2000) for its assumptions about the existence of {\\it ad hoc} mechanisms that regulate the accretion flow on to the WD (e.g. Hachisu et al. 1996). Observationally, Badenes et al. (2007) have noted the absence, in seven nearby SN-Ia remnants, of the signatures of the strong wind from the accretor that supposedly stabilizes the accretion flow, and permits reaching $M_{\\rm ch}$. Prieto et al. (2007) have found no evidence for a low-metallicity threshold in SN-Ia hosts, in contrast to the predictions by Kobayashi et al. (1998; see also Kobayashi \\& Nomoto 2008) of such a threshold, due to a minimum metallicity that is required for the wind regulation mechanism to be effective. While evidence for circumstellar material, consistent with expectations from a wind from a red-giant companion, has been found in one recent normal SN-Ia (Patat et al. 2007), such material is not observed in other events (Mattila et al. 2005; Simon et al. 2007; Leonard 2007). For the remnant of Tycho's SN, which was a type-Ia (Badenes et al. 2006), there have been conflicting claims about the identification and nature of a remaining companion star (Ruiz-Lapuente et al. 2004; Fuhrmann 2005; Ihara et al. 2007). If, in the end, no companion is found, this would be another problem for the SD picture. Recent metal abundance measurements suggest a type-Ia explosion also in Kepler's SN remnant (e.g., Blair et al. 2007; Reynolds et al. 2007) but the velocity of the remnant away from the Galactic plane, together with evidence for a circumstellar medium from a massive progenitor, are problematic the context of a binary progenitor system (Reynolds et al. 2007). For core-collapse SNe, which derive from massive ($\\ga 8 M_{\\odot}$) stars, direct searches of pre-explosion images for progenitors have had a number of successes -- in the case of SN1987A (West et al. 1987; White \\& Malin 1987), and several other events with suitable {\\it Hubble Space Telescope} (HST) data (Maund \\& Smartt 2005; Li et al. 2006; Hendry et al. 2006; Gal-Yam et al. 2007). In contrast, the expectation, based on the SD and DD models, that SNe-Ia explode in old, low-mass, systems, has discouraged such direct searches for progenitors of SNe-Ia in the HST era. While there have been some studies of SN-Ia environments using post-explosion HST data (Van Dyk et al. 1999), the reverse has not been done, largely due to a preconception that the progenitors would be undetectable. (Interestingly, Van Dyk et al. 1999 noted young stellar populations in the vicinity of the four SNe-Ia they studied, but subtracted them off the images, arguing that the SNe-Ia were necessarily derived from an old population). However, several recent observational developments, in addition to the problems faced by the popular models, suggest that such a search may be useful after all. Mannucci et al. (2005) have measured stellar-mass-normalized type-Ia SN rates as a function of galaxy Hubble type and galaxy colors, and have found that the SN-Ia rate in star-forming galaxies traces the star formation rate. Such a correlation directly implies that at least some of the progenitors are young stars. This study and several subsequent ones (Scannapieco \\& Bildsten 2005; Mannucci et al. 2006; Sullivan et al. 2006; Dilday et al. 2008) have demonstrated that there must be a wide distribution of delay times between star formation and SN explosion, as predicted by some progenitor models (e.g., Greggio \\& Renzini, 1983; Greggio 2005). The spread in the delay times can be described in terms of two populations of progenitors: a ``prompt'' one, which dominates the SN-Ia rate in star-forming galaxies and has a typical delay times of less than $\\sim 10^8$~yr; and ``tardy'' population, having a SN-Ia delay time of $\\ga 10^9$~yr (or a large-delay tail to the same progenitor population above), needed to produce also the SN-Ia rate measured in old stellar populations with no current star formation. An outstanding puzzle is the possible decrease in the SN-Ia rate at redshifts $z>1$ measured by Dahlen et al. (2004, 2008), at a cosmic time when the star-formation rate was an order of magnitude higher than today. The SN-Ia rate, which would therefore have then been dominated by the prompt population, would be expected to track the star formation rate, as the latter flattens or continues to rise to high $z$. On the other hand, reanalysis of some of these data by Kuznetsova et al. (2008), and independent high-$z$ rates by Poznanski et al. (2007), have questioned the significance of the claimed decrease in SN-Ia rate at high redshift, pointing to the need for further observations to resolve the issue. In any event, the formation of a degenerate carbon stellar core within $10^8$~yr for a significant fraction of SN-Ia progenitors points to stars with zero-age main sequence (ZAMS) masses of $\\ga 5 M_\\odot$ (e.g. Girardi et al. 2000). During their post-main-sequence evolution, such intermediate-mass stars may reach luminosities high enough to make them potentially detectable in pre-explosion HST images of nearby host galaxies. Such evolved stars could be present in SN-Ia progenitor systems, either in the role of mass donors to the WD (previously produced by a more massive primary), where the evolutionary timescale of the secondary dictates the onset of the explosion; or in the role of single-star SN-Ia progenitors, where degenerate carbon-core ignition occurs in an evolved star that has somehow lost its hydrogen envelope (e.g. Tout 2005; Waldman, Yungelson, \\& Barkat 2007). Furthermore, recent measurements of extragalactic SN-Ia rates at various redshifts have yielded high rates, especially in star-forming environments. Maoz (2008) has recently compiled, and compared self-consistently, different measurements of various observables related to SN rates, and used them to estimate the fraction of intermediate-mass close binaries that explode as SNe-Ia through the prompt and tardy channels. He shows that the high SN-Ia rates (as well as other independent observables such as intracluster abundances) indicate that most or all close, intermediate-mass, binaries must explode as SNe-Ia, with, possibly, an excess of SNe-Ia over progenitor systems. This conclusion holds despite the uncertainties in the initial mass ranges that lead to viable WD SN-Ia progenitors in the SD and DD scenarios, and the uncertainties in the initial parameters of binaries (binarity fraction, mass ratio distributions, separation distributions). The conclusion stands in contrast to detailed SD and DD models that predict a small exploding fraction among the intermediate-mass population. The independent observation that a major part of the SN-Ia population derives from a young population, indicating initial masses in a narrow range of $\\sim 5-8 M_\\odot$ (see above), further culls the pool of potential progenitors, and makes the case for an excess of SN-Ia explosions compared to progenitors almost unavoidable. A possible solution could, again, be evolved, single-star, stripped envelope, SN-Ia progenitors. Direct searches for such stars in pre-explosion images of the sites of SNe-Ia in nearby galaxies can test this scenario. In this paper, we perform such a search in the massive nearby galaxy NGC~1316, a prolific SN producer that hosted two SN-Ia events in 2006, by analyzing deep pre-explosion HST images. After compiling some facts about this galaxy and its SNe, we describe the archival HST and ground-based data we have used, and their analysis for setting flux limits on individual progenitors for each of the two SNe. We then discuss the physical limits on SN-Ia progenitors imposed by these results. ", "conclusions": "In this section, we now examine, based on stellar evolution models, which stellar progenitors present 3 years prior to the explosion can be ruled out by our limits. Metallicity that is close to Solar has been measured in this galaxy, both for the stars Goudfrooij et al. (2001a,b) and for the gas Kim \\& Fabbiano (2003). \\begin{figure*} \\includegraphics[width=0.85\\textwidth]{n1316fig3color.ps} \\caption{Model stellar evolution tracks from ZAMS to AGB (solid curves, from Girardi et al. 2000, for $3-6M_\\odot$, and from Salasnich et al. 2000, for $8-12 M_\\odot$), for the TP-AGB phase (Marigo \\& Girardi 2007; dotted curves), and for post-AGB evolution (Bloecker 1995b; dashed curves), shown in the plane of absolute $I$-band AB magnitude, $M_I$, versus $V-I$ and $B-I$ AB colours. Models are labeled according to their ZAMS masses, in Solar units, with smaller labels for the post-AGB tracks. The $3M_\\odot$ post-AGB model is one which, among the models presented by Bloecker (1995b), has a post-AGB mass of $0.625 M_\\odot$. The grey areas show the regions of parameter space with luminosities greater than our absolute magnitude limits in $B$, $V$, and $I$, for stars at the positions of the SNe in NGC~1316 (assuming a distance of 19~Mpc and no extinction).} \\label{tracks} \\end{figure*} Figure~\\ref{tracks} shows, in a color-absolute-magnitude diagram, various Solar-metallicity stellar evolution tracks. In the range $3-6 M_\\odot$ ZAMS mass, the tracks are from Girardi et al. (2000), from the main sequence to the onset of thermal pulsations on the AGB branch. The models shown include convective overshooting. Recent models by Marigo \\& Girardi (2007) for the thermal-pulsation stage on the AGB are shown for $3-5 M_\\odot$ ZAMS masses. For $8-12 M_\\odot$ stars, the models are by Salasnich et al. (2000) between the ZAMS and carbon core ignition, assuming Solar composition. The dashed curves show post-AGB tracks by Bloecker (1995b) for stars of ZAMS masses $3-7 M_\\odot$. Bolometric luminosities and effective temperatures of the models were converted to broad-band AB magnitudes by integrating the surface fluxes in stellar-atmosphere model spectra over the appropriate bandpasses. The solar-metallicity stellar atmospheres for a grid of temperatures and surface gravities were interpolated to the temperatures and gravities of the stellar evolution tracks. The stellar atmospheres are from the BaSeL 2.2 database (Lejeuene et al. 1997, 1998; Westera et al. 2002), and from Kurucz (1992) for temperatures above $4000 K$. For the post-AGB models, monochromatic luminosities were obtained assuming a blackbody spectrum, an approximation which is accurate to a few percent for these hot stars and the optical bands considered here. Models are labeled according to their ZAMS masses, in Solar units. The grey-shaded areas show the regions of parameter space with luminosities greater than our absolute magnitude limits in $B$, $V$, and $I$, assuming no extinction, as listed in Table~1. Adoption of the plausible extinctions listed in Table~1 would shift the grey region upward by 0.1~mag, and to the left by 0.15~mag (left panel) and 0.3~mag (right panel). We see that our upper limits on the stellar luminosities at the sites of the two SNe, do probe some of the parameter space, even if small, of the late-time evolution of intermediate-mass stars. Specifically, according to the models shown, stars with ZAMS masses $\\ga 6M_\\odot$ reach luminosities higher than our limits at the end of their AGB phases. Stars with initial masses $\\ga 9 M_\\odot$ spend much of the time from the first ascent to the red-giant branch and onward at luminosities above the limits. Similarly, post-AGB stars of ZAMS masses $\\ga 4 M_\\odot$, at the beginning of their constant luminosity path to high temperatures on the HR diagram (which results in a monotonic decrease in optical luminosity, see Fig.~\\ref{tracks}), may be above our upper limits. However, the uncertainties that affect models of the later stages of stellar evolution, particularly uncertainties due to convective overshooting and the details of mixing and mass loss (e.g., Herwig 2005), make the conclusions based on comparison to such models tentative. Indeed, Pauldrach et al. (2004), when modeling the ultraviolet spectra of nine central stars of planetary nebulae (CSPNs), have found severe discrepancies between the masses deduced for five of the stars, compared with those predicted by the mass-luminosity relation from theoretical post-AGB stellar models. Remarkably, the masses of these stars are close to the Chandrasekhar limit. Conversely, Napiwotzki (2006) and Gesicki \\& Zijlstra (2007) have concluded, based, on different analysis methods, that CSPNs, including some of those analyzed by Pauldrach et al. (2004), have masses of $\\sim 0.6 M_\\odot$, similar to most WDs, and with mass-luminosity relations consistent with post-AGB model predictions. In any case, in standard SD models of SNe-Ia, complex binary evolution precedes a SN-Ia explosion, and hence using single-star models to predict the colours and the luminosities of the donor stars could be inaccurate. And, if SNe-Ia evolve from single stars within some initial mass and metallicity ranges, which have somehow (e.g. via mass loss, or previous binary interactions) lost their hydrogen envelopes, then the stellar tracks for normal AGB evolution that we have used are unlikely to provide reliable predictions for the properties of such progenitors. It should also be kept in mind that AGB stars, which are prodigious dust producers during their thermal-pulse stages, are likely enshrouded in dust shells at the end of their evolution, and the exact point at which they become visible again on the post-AGN branch is uncertain (e.g., Bloecker 1995a,b). In summary, we have searched pre-explosion images of the sites of two SNe-Ia for potential luminous stellar progenitors, with null results. As always is the case in such studies, our null result does not conclusively rule out any of the scenarios or mass ranges for SN-Ia progenitors. Under any of the possibile scenarios of strong obscuration by dust (by unresolved clumps, or by a screen that is behind the bulk of the stellar mass of the galaxy, or by a circumstellar shell), more luminous progenitors could be hidden. Furthermore, model predictions for the later stages of stellar evolution are still quite uncertain, all the more so if binary interactions take place. Finally, SNe 2006dd and 2006mr may have, by chance, belonged to the tardy SN-Ia population, in which case one would not expect them to be associated with a young population. Nevertheless, our study shows that deep HST observations of nearby SN hosts can graze some interesting regions of stellar parameter space that may be relevant for young-population SN-Ia progenitors. Future SNe-Ia will explode also in more nearby galaxies that have HST imagery. This will improve the statistics of such searches, will make accessible lower stellar luminosities, and will perhaps eventually reveal an actual progenitor. In view of the many unknowns behind SN-Ia formation, such observational studies may provide valuable clues." }, "0801/0801.2917_arXiv.txt": { "abstract": "We report limits in the planetary-mass regime for companions around the nearest single white dwarf to the Sun, van Maanen's star (vMa\\,2), from deep $J$-band imaging with Gemini North and {\\it Spitzer} IRAC mid-IR photometry. We find no resolved common proper motion companions to vMa\\,2 at separations from $3 - 45\\arcsec$, at a limiting magnitude of $J\\approx23$. Assuming a total age for the system of $4.1\\pm1$\\,Gyr, and utilising the latest evolutionary models for substellar objects, this limit is equivalent to companion masses $>7\\pm1\\,\\mjup$ ($T_{\\rm eff}\\approx300$\\,K). Taking into account the likely orbital evolution of very low mass companions in the post-main sequence phase, these $J$-band observations effectively survey orbits around the white dwarf {\\it progenitor} from $3 - 50$\\,AU. There is no flux excess detected in any of the complimentary {\\it Spitzer} IRAC mid-IR filters We fit a DZ white dwarf model atmosphere to the optical $BVRI$, 2MASS $JHK$ and IRAC photometry. The best solution gives $T_{\\rm eff}=6030\\pm240$K, log~g$=8.10\\pm0.04$ and, hence, $M= 0.633\\pm0.022\\Msun$. We then place a $3\\,\\sigma$ upper limit of $10\\pm2\\,\\mjup$ on the mass of any unresolved companion in the $4.5\\mu$m band. ", "introduction": "Direct imaging of extra-solar planetary-mass companions to solar-type stars is complicated by problems of contrast and resolution. At the time of writing, no planet has been directly imaged around a solar-type star. An alternative solution to these difficulties is to target intrinsically faint stars instead, such as white dwarfs. Stellar evolution lends two huge advantages when searching for very faint companions: white dwarfs are up to $\\sim10^4$ times fainter than their main sequence progenitors, and the orbits of any planetary-mass companions that lie outside the stellar envelopes during the giant phases will expand outwards as mass is lost from the central star, increasing the projected separation by a maximum factor $M_{\\rm MS} / M_{\\rm WD}$ \\citep{jeans}. Thus, the problems of contrast and resolution are greatly reduced. The possible evolution of planetary systems in the post-main sequence phase is discussed in more detail by \\citet{duncan}, \\citet*{burleigh02}, \\citet{Debes02} and \\citet{villaver}. The direct detection of such low mass companions to white dwarfs opens up the possibility for spectroscopic investigation of a previously unobserved class of object: evolved low-mass brown dwarfs and planetary-mass gas giants as low in temperature as $T_{\\rm eff}\\approx300$\\,K. In contrast, the directly imaged planetary-mass ($\\approx5\\mjup$) companion to the brown dwarf 2MASSW\\,J1207 has the spectrum of a mid-L~dwarf \\citep{chauvina,chauvinb}, because it is still young ($\\sim10^7$~years old). The coolest known brown dwarf, ULAS\\,J$003402.77-005206.7$, has a temperature $600$K$< T_{\\rm eff} < 700$K \\citep{warren07} and a spectral type T\\,8.5. The letter ``Y'' has been suggested for the next, cooler, spectral type \\citep{Kirkpatrick}, which might include evolved planetary-mass companions to white dwarfs. Alternatively, if no obvious spectral change triggers the use of a new letter, then the T classification will need to be extended beyond T\\,8.5. The idea of using white dwarfs to find intrinsically faint, low mass companions is not new. \\citet{probst83} and \\citet{becklin88} used the low luminosity of white dwarfs to search for brown dwarf companions as near-infrared photometric excesses, and indeed the latter achieved success with GD\\,165\\,B (L\\,4). More recently, \\citet{fbz05} have conducted a comprehensive search for brown dwarf companions to several hundred white dwarfs but detected only one new pair (GD\\,1400, L\\,$6-7$, \\citealt{GD1400}, \\citealt{dobbie05}), while \\citet{maxted06} and \\citet{wd0137b} have detected a close L\\,8 companion to the white dwarf WD\\,0137$-$349. No other detached substellar companions to white dwarfs are known. The most intriguing circumstantial evidence for the existence of old planetary systems that have survived to the final stage of stellar evolution comes from the discovery of metal-rich circumstellar dust and gas disks around a growing number of white dwarfs \\citep{zuckerman87,becklin05,gd362,gd56,sdss1228,vh07, Jura07,WD1150}. {\\it Spitzer} mid-infrared spectroscopy has now revealed that the dust disks are composed largely of silicates \\citep{Reach,JuraGD362}. The favoured explanation for the origin of this material is the tidal disruption of an asteroid that has wandered into the Roche radius of the white dwarf \\citep{Jura1}, perhaps through interaction with planets in a solar system that has become destabilized in the wake of the planetary nebula phase \\citep{Debes02}. \\citet{Jura2} suggests that at least 7\\% of white dwarfs possess asteroid belts. If that is indeed the case, it is likely that at least this number of white dwarfs also possess planetary systems. The recent discovery of a $M {\\rm sin} i = 3.2\\mjup$ planet in a $1.7$\\,AU orbit around the extreme horizontal branch star V391\\,Pegasi by \\citet{V391Peg} proves that such objects can survive red giant branch evolution, strongly suggesting that it is simply a matter of time before a planet is discovered around a white dwarf. \\citet{burleigh02} made predictions concerning the likely near-infrared brightness of putative resolved, giant planetary companions to nearby white dwarfs, based on their likely total ages and distances. In 2002 we initiated a programme to search for wide, spatially resolved very low mass common proper motion companions to white dwarfs via direct imaging. In particular, we aim to find substellar companions with masses greater than a few~$\\,\\mjup$ around white dwarfs within $\\approx20$~pc of the Sun. Such companions are expected to have near-IR magnitudes brighter than $J\\sim23.5$, commensurate with the expected sensitivity of an 8m telescope in a one hour exposure. \\begin{sloppypar} We christened our project {\\it ``DODO''} -- {\\it D}egenerate {\\it O}bjects around {\\it D}egenerate {\\it O}bjects. Preliminary results and progress reports have been published elsewhere \\citep*{dodoa,dodob,dodoc}. Here, we report our results for the nearest single white dwarf to the Sun in our sample, van Maanen's star (vMa\\,2, WD\\,0046$+$051, d$\\,= 4.41$\\,pc, \\citealt{Hipparcos}). We combine two epochs of deep, ground-based $J$-band imaging from the 8m Gemini North telescope with {\\it Spitzer} mid-infrared photometry to place limits on the masses and temperatures of any common proper motion and unresolved ultra-cool substellar and planetary-mass companions. \\end{sloppypar} \\subsection{vMa\\,2 and previous infrared observations} vMa\\,2 was discovered serendipitously by \\citet{vMa2} in a survey for common proper motion companions to an unrelated star, HD\\,4628. It has a cool, helium-rich atmosphere and strong resonance Ca\\,II H \\& K lines in its optical spectrum, and is classified as a DZ white dwarf. The heavy elements are expected to sink below the photosphere on a timescale much shorter than the white dwarf cooling time. Their presence has been explained previously in terms of episodic accretion from the interstellar medium \\citep{Dupuis92}, but this scenario requires the accretion rate of hydrogen to be at least two orders of magnitude lower than the metals (\\citealt{Wolff02, Dupuis93, Dufour}). Alternatively, the presence of metals in DZs might be explained from accretion of cometary material or tidally disrupted asteroids and planets. However, no DZ has been identified with an infrared excess due to dust emission, possibly because the diffusion timescale for heavy elements in cool white dwarf helium-rich atmospheres is long and the material may have been accreted as much as $\\sim10^6$~years ago. Nonetheless, DZ white dwarfs are a speculative candidate for hosting old planetary systems and vMa\\,2, as the nearest single white dwarf to the Sun, is an ideal target for such a search. Through analysis of {\\it Hipparcos} data, \\citet{Makarov} claimed to have detected astrometrically a $0.06 \\pm 0.02\\msun$ substellar companion to vMa\\,2, with an orbital period of 1.57~years and a maximum separation on the sky of $0.3\\arcsec$. \\citet*{fbm} carried out a search for this companion by direct imaging with adaptive optics in the mid-infrared $L'$ band. They also looked for an unresolved companion as a near- and mid-infrared photometric excess in ground-based and {\\it ISO} photometry. They refuted the existence of Makarov's companion, and of any excess emission due to dust, and placed a limit of $T_{\\rm eff} \\la 500$\\,K on the temperature of any putative substellar companion. \\begin{table*} \\caption{Adopted parameters for van Maanen's star} \\begin{center} \\begin{tabular}{cccccccccc} \\hline $\\mu\\,^a$ & $\\theta\\,^a$ & d$\\,^a$ & $T_{\\rm eff}\\,^b$ & log~$g\\,^b$ & $M_{\\rm WD}\\,^b$ & $t_{\\rm WD}\\,^c$ & $M_{\\rm MS}\\,^d$ & $t_{\\rm MS}\\,^e$ & $t_{\\rm total}$ \\\\ (mas/yr) & (mas/yr) & (pc) & (K) & & ($\\msun$) & (Gyr) & ($\\msun$) & (Gyr) & (Gyr) \\\\ \\hline \\hline 1231.72 & $-2707.67$ & 4.41 & 6030 (240) & 8.10 (0.04) & 0.633 (0.022) & 3.17 (0.29) & 2.6 & 0.9 & 4.1 \\\\ \\hline \\end{tabular} \\end{center} $^a$~Hipparcos measurements, $^b$~from our new fit to the optical, near-IR and IRAC mid-IR photometry (see section~4.2), $^c$~estimated using evolutionary models appropriate for He-atmosphere white dwarfs with C/O core compositions, see http://www.astro.umontreal.ca/~bergeron/CollingModels/, $^d$~derived from the initial-final mass relation of \\citet{Dobbie06}, $^e$~\\citet{Wood} \\end{table*} ", "conclusions": "We have placed limits on ultra-cool substellar and planetary-mass objects around the nearest single white dwarf, vMa\\,2, through a search for common proper motion companions and mid-infrared photometric excesses. The red giant progenitor to vMa\\,2 had a maximum radius of $\\sim1000R_\\odot$ ($4.6$\\,AU, \\citealt*{Hurley}). Therefore, taking into account the post main sequence evolution of the orbits of any companions, our $J$-band images have covered all orbits of substellar objects that originally lay beyond the maximum extent of the red giant envelope, up to 50\\,AU distance. The complimentary {\\it Spitzer} IRAC photometry places slightly higher limits on the masses and temperatures of any spatially unresolved substellar companion whose orbit fortuitously lies along the line of sight to vMa\\,2, or that currently has an orbit within $\\approx13$\\,AU of the white dwarf. These limits are significantly lower than those reported by previous studies, e.g. \\citet{fbm} and \\citet{fbz05}. A variety of searches for planetary companions to white dwarfs are currently underway (e.g.~\\citealt{dodob}; \\citealt{Debes1}; \\citealt{MullallyB}). Further, extensive results from the {\\it DODO} survey are in preparation \\citep{Hogan}." }, "0801/0801.0409_arXiv.txt": { "abstract": "{ Simbol-X is a French-Italian mission, with a participation of German laboratories, for X-ray astronomy in the wide 0.5-80 keV band. Taking advantage of emerging technology in mirror manufacturing and spacecraft formation flying, Simbol-X will push grazing incidence imaging up to $\\sim80$ keV, providing an improvement of roughly three orders of magnitude in sensitivity and angular resolution compared to all instruments that have operated so far above 10 keV. This will open a new window in X-ray astronomy, allowing breakthrough studies on black hole physics and census and particle acceleration mechanisms. We describe briefly the main scientific goals of the Simbol-X mission, giving a few examples aimed at highlighting key issues of the Simbol-X design. ", "introduction": "A seminal result obtained with HEAO-1 at the end of the 70' is the precise measure of the spectrum (from a few keV up to $\\sim100$ keV) of the isotropic, extragalactic Cosmic X-Ray Background (CXB), discovered by Riccardo Giacconi, Bruno Rossi and collaborators during one of the first rocket-borne X--ray experiments in 1962. The HEAO1 data showed that the CXB energy density has a broad maximum around 30 keV, where it is about 5 times higher than at 1 keV and 50\\% higher than at 10 keV. It was soon realized that the CXB is most likely due to the contribution of many discrete sources at cosmological distances \\citep{sw79}. Most of these sources are active galactic nuclei, AGN, implying that the CXB energy density provides an integral estimate of the mass accretion rate in the Universe, and therefore of the super-massive black hole (SMBH) growth and mass density. Unfortunately, the integrated light from all sources detected in HEAO1 all-sky survey could directly explain only less than 1\\% of the CXB. Indeed, the use of collimated detectors on board first UHURU and Ariel-V, and then HEAO1 in the 1970 decade led to the discovery of $<1000$ X-ray sources in the whole sky. X-ray imaging observations, performed first by {\\it Einstein} and ROSAT in the soft X-ray band below $\\sim3$ keV, and then by ASCA, BeppoSAX, XMM--Newton and Chandra up to 8-10 keV, detected tens of thousands of X-ray sources, and resolved nearly 100\\% of the CXB below a few keV and up to 50\\% at 6-8 keV. These observations increased by orders of magnitude the discovery space for compact objects (both Galactic neutron stars and black holes and AGN) and for thermal plasma sources. However, they still leave open fundamental issues, such as what is making most of the energy density of the CXB at $\\sim30$ keV. Above 10 keV the most sensitive observations have been performed so far by collimated instruments, like the BeppoSAX PDS, and by coded masks instruments, like INTEGRAL IBIS and Swift BAT. Only a few hundred sources are know in the whole sky in the 10-100 keV band, a situation recalling the pre--{\\it Einstein} era at software energies. A new window in X-ray astronomy above 10 keV must be opened, producing an increase of the discovery space similar to that obtained with the first X-ray imaging missions. This will be achieved by Simbol-X, a formation flight mission currently under preparation by France and Italy, with a participation from German laboratories. Very much like the {\\it Einstein} Observatory, this mission will have the capabilities to investigate almost any type of X-ray source, from Galactic and extragalactic compact sources, supernova remnant (SNR), young stellar objects and clusters of galaxies, right in the domain where accretion processes and acceleration mechanisms have their main signatures. This paper summarizes the main scientific goals of the Simbol-X mission, putting the emphasis on the core science objectives. ", "conclusions": "Thanks to the emerging technology in X-ray mirrors (e.g. multilayer coating, see Pareschi et al. these proceedings) and spacecraft formation flying, Simbol-X will provide a large collecting area (of the order of 100-1000 cm$^{-2}$) from a fraction of keV up to $\\sim80$ keV, thus overcoming the ``10 keV limit'' for high accuracy imaging and spectroscopy of all past and current X-ray observatories. This, together to the good image quality (PSF HPD$<20$ arcsec, FWHM$<10$ arcsec), relatively large field of view (12 arcmin diameter), good detector quantum efficiency, resolution and low internal background, will allow breakthrough studies on black hole physics and census, and particle acceleration mechanisms." }, "0801/0801.1914_arXiv.txt": { "abstract": "We give a closer look at the Central Limit Theorem (CLT) behavior in quasi-stationary states of the Hamiltonian Mean Field model, a paradigmatic one for long-range-interacting classical many-body systems. We present new calculations which show that, following their time evolution, we can observe and classify three kinds of long-standing quasi-stationary states (QSS) with different correlations. The frequency of occurrence of each class depends on the size of the system. The different microsocopic nature of the QSS leads to different dynamical correlations and therefore to different results for the observed CLT behavior. ", "introduction": "Very recently there has been a lot of interest in generalizations of the Central Limit Theorem (CLT) \\cite{clt0,umarov,barkai} and on their possible (strict or numerically approximate) application to systems with long range correlations \\cite{luis,hil}, at the edge-of-chaos \\cite{ucc}, nonlinear dynamical systems the maximal Lyapunov exponent of which is either exactly zero or tends to vanish in the thermodynamic limit (increasingly large systems) \\cite{anteneodo}, hindering this way mixing and thus the application of standard statistical mechanics. A possible application of nonextensive statistical mechanics \\cite{cost1,clt1} has been advocated in these cases. Along this line we discuss in the present paper a detailed study of a paradigmatic \\textit{toy model} for long-range interacting Hamiltonian systems \\cite{hmf,liap,epn-rap,chavanis,fulvio1,ruffo,reply07}, i.e. the Hamiltonian Mean Field (HMF) model which has been intensively studied in the last years. In a recent article \\cite{rapis-ctnext07}, we presented molecular dynamics numerical results for the HMF model showing three kinds of quasi-stationary states (QSS) starting from the same water-bag initial condition with unitary magnetization ($M_0=1$). In the following we present how the applicability of the standard or $q$-generalized CLT is influenced by the different microscopic dynamics observed in the three classes of QSS. In general, averaging over the threee classes can be misleading. Indeed, the frequency of appearence of each of these classes depends on the size of the system under investigation, and there is no clear evidence that a predominant class exists. ", "conclusions": "On the basis of the new calculations presented here, one should distinguish among different classes of QSS for a given size of the HMF system. Our tests do confirm that correlations can be different for different dynamical realizations of the same system, starting from the same class of initial conditions, and therefore also the Central Limit Theorem behavior can change. According to the class considered we can have a Gaussian Pdf, a $q$-Gaussian one or a mixture between the two. In this respect, the presence of a $q$-Gaussian curve confirms the results previously published about the indications for a generalized CLT in long-range Hamiltonian systems when ergodicity is broken and correlations are strong enough. The present results, where we verify that the time evolution of the system is quite sensitive to the specific initial conditions, reminds what occurs in other anomalous systems, like the Hydrogen atom as studied recently in Refs. \\cite{Oliveira1,Oliveira2}. We have checked also that inequivalence between time averages and ensemble averages continues to hold within the different classes for $M_0=1$ and for $M_0=0$ initial conditions. Finally, although it is true that the frequency of occurrence of events of class 1 and 2 for very large sizes, and in particular the attractor for the events of class 2, remain to be investigated with more accuracy (calculations with better statistics are in progress) before advancing definitive conclusions, from these simulations one could conjecture that the events of class 1 tend to disappear in the thermodynamic limit. However, even if this was the case, their probability of occurrence (as well as the probability of occurrence of the first metastable plateau of events of class 2) can be significantly different from zero for large but finite systems and therefore the applicability of $q$-statistics in metastable states of real complex systems remains a valid and interesting possibility." }, "0801/0801.3857_arXiv.txt": { "abstract": "We present evidence of Fe fluorescent emission in the \\cha\\ \\hetgs\\ spectrum of the single G-type giant \\hr\\ during a large flare. In analogy to solar X-ray observations, we interpret the observed Fe~K$\\alpha$ line as being produced by illumination of the photosphere by ionizing coronal X-rays, in which case, for a given Fe photospheric abundance, its intensity depends on the height of the X-ray source. The \\hetgs\\ observations, together with 3D Monte Carlo calculations to model the fluorescence emission, are used to obtain a direct geometric constraint on the scale height of the flaring coronal plasma. We compute the Fe fluorescent emission induced by the emission of a single flaring coronal loop which well reproduces the observed X-ray temporal and spectral properties according to a detailed hydrodynamic modeling. The predicted Fe fluorescent emission is in good agreement with the observed value within observational uncertainties, pointing to a scale height $\\lesssim 0.3$\\rstar. Comparison of the \\hr\\ flare with that recently observed on II~Peg by {\\em Swift} indicates the latter is consistent with excitation by X-ray photoionization. ", "introduction": "\\label{s:intro} Since the early 1970's, spatially resolved observations of the solar corona have revealed a high degree of structuring over all scales, from of order of the solar radius down to instrumental resolution limits (e.g., \\citealt{Vaiana73,Vaiana73b}). This spatial structuring is intimately linked to the characteristics of magnetic field generation and interaction, and to plasma heating mechanisms. On other stars, especially very active ones (with X-ray luminosity $L_{\\rm X}$ up to $10^4 \\times L_{\\rm X \\odot}$), coronal structure and its relation with stellar parameters such as mass, rotation and evolutionary phase, remains very uncertain. Techniques used to date to investigate the morphology of stellar coronae comprise rotational modulation \\citep[e.g.,][]{MarinoL03,Flaccomio05}, flare modeling (e.g., \\citealt{Reale04}; \\citealt{Testa07a}, hereafter \\ta), eclipse mapping \\citep[e.g.,][]{White90}, spectroscopic density and radiation field diagnostics \\citep[e.g.,][]{Testa04b,Ness04}, resonance scattering \\citep[e.g.,][]{Testa04a,Matranga05,Testa07b}, velocity modulation \\citep[e.g.,][]{Brickhouse01,Chung04,Hussain05,Huenemoerder06}, and simultaneous Doppler imaging and X-ray spectroscopy \\citep{Hussain07}. The fluorescent iron line at $\\sim$6.4~keV \\footnote{The Fe fluorescent feature actually consists of two components, at 6.391 and 6.404~keV (for Fe\\,{\\sc i}), which are unresolved by present instruments, therefore hereafter we will refer to the feature as to a single emission line.} presents a further, potentially powerful diagnostic of stellar coronal geometry. Extensively used in study of active galactic nuclei and X-ray binaries (see e.g.\\ reviews \\citealt{George91,Reynolds03}), and often seen on the Sun during flares \\citep{Culhane81,Parmar84,Tanaka84,Zarro92,Phillips94}, the line is produced by electron cascade after one of the two K-shell ($n$=1) electrons of an iron atom (or ion) is ejected following photoelectric absorption of X-rays. In the solar and stellar case, for a given irradiating flare spectrum the line strength depends essentially on three parameters: the flare height, heliocentric angle, and photospheric Fe abundance (e.g., \\citealt{Bai79}; \\citealt{Drake07b}, hereafter D07b). If it can be observed in stars, Fe~K fluorescence therefore presents a means of estimating flare and coronal scale height. Fe fluorescence has now been detected on a number of pre-main sequence stars with disks \\citep[e.g.,][]{Tsujimoto05,Favata05}, where it is most likely produced by coronal X-ray irradiation of the cold disk material. Here we present evidence for {\\em photospheric} Fe fluorescent emission in the \\cha\\ \\hetgs\\ spectrum of the X-ray active single G1 giant \\hr\\ (see \\ta\\ and references therein for a description of the characteristics of the target). The observations and analysis are briefly described in \\S\\ref{s:obs}. In \\S\\ref{s:model} we present 3D Monte Carlo calculations of the Fe K$\\alpha$ fluorescence. The results are presented in \\S\\ref{s:results}, where we derive an estimate for the coronal scale height and compare it with the prediction of the loop hydrodynamic model that succeeds in describing the observed flaring spectrum (\\ta). To our knowledge, the \\hr\\ {\\em Chandra} spectrum represents only the second detection of Fe photospheric fluorescence. Very recently, the line was observed by the {\\em Swift} X-ray Telescope during a ``superflare'' on the RS~CVn binary system II~Peg by \\citet{Osten07}, who ascribed the excitation mechanism to electron impact ionization of photospheric Fe instead of photoionization. We discuss our result and the II~Peg observations in \\S\\ref{s:discuss}, and show that the latter is also consistent with, and more plausibly associated with, photoionization than electron impact. ", "conclusions": "\\label{s:discuss} We have presented the first detailed analysis of observed photospheric fluorescence to probe the geometric properties of stellar coronal emission. For the flare on \\hr\\ we find the observed Fe~K$\\alpha$ intensity completely consistent with production through photospheric irradiation by flare X-rays. Our estimate for the flare scale height of $h \\lesssim 0.3$\\rstar\\ provides a cross-check for the results of the hydrodynamic modeling of the flare observed on \\hr\\ that can be satisfactorily reproduced by a single flaring loop of semi-length $L =0.5$\\rstar. The agreement between the observed $\\mathcal{E}$ and the value computed directly from this model provides further confidence both in the hydrodynamic modeling approach and X-ray fluorescence as the excitation mechanism for the Fe~K$\\alpha$ line. We note that our model based on \\cha\\ data might underestimate the high energy continuum flux because we cannot detect any additional (either thermal or non-thermal) emission possibly present at energies outside the \\cha\\ band. However, if additional high energy emission were present, the higher flux above 7.11~keV would make the efficiency lower bringing it in even better agreement with the predictions of the model. As noted in \\S\\ref{s:intro}, Fe K$\\alpha$ emission has been extensively observed in solar X-ray spectra during flares \\citep{Culhane81,Parmar84,Tanaka84,Zarro92,Phillips94}. Two main production mechanisms were initially explored to explain the line: electron impact from a non-thermal electron population, and X-ray photoionization. The observed Fe K$\\alpha$ evolution with respect to the thermal and non-thermal emission, and the observed center-to-limb variations seen on the Sun, strongly support the X-ray fluorescence mechanism \\citep{Culhane81,Parmar84,Tanaka84}, although in rare cases there is indication that electron impact might contribute during the early impulsive phase \\citep[e.g.,][]{Emslie86}. \\citet{Ballantyne03} have also shown that Fe~K production in accretion disks by non-thermal electron bombardment is extremely unlikely, and requires 2-4 orders of magnitude greater energy dissipation in the electron beam than is required for an X-ray photoionization source. In this context, the recent observations of an extremely large Fe~K$\\alpha$ equivalent width in the PMS star V1486~Ori \\citep{Czesla07}, and variability in Fe~K$\\alpha$ emission uncorrelated with the observed X-ray continuum on Elias 29 \\citep{Giardino07}, seem to challenge the photoionization excitation mechanism. However, D07b note that fluorescence from PMS disks excited by X-rays originating from the unseen stellar hemisphere will also be observed: in $\\leq 50$\\%\\ of cases, Fe~K$\\alpha$ can then appear anomalous when compared only to the {\\em observed} thermal continuum. To date, the {\\em Swift} observation of II~Peg \\citep{Osten07} is the only other case where Fe~K$\\alpha$ emission has been observed for a star (other than the Sun) lacking substantial circumstellar material. \\cite{Osten07} dismissed X-ray fluorescence as the excitation mechanism based on arguments for fluorescence in an optically-thin medium, and instead attributed the Fe K$\\alpha$ to electron impact with non-thermal electrons. As noted in earlier discussions \\citep[e.g.][]{Parmar84,Ballantyne03}, one major problem for an electron impact excitation mechanism is the inefficiency of this process. Extremely large energies must be dissipated in accelerated electron beams to produce continuum and fluorescence emission that might be observed on a star. Indeed, \\citet{Osten07} showed that the hard X-ray continuum observed in the II~Peg flare would require an energy of $3 \\times 10^{40}$~ergs in non-thermal electrons if interpreted in terms of thick-target bremsstrahlung. Over the time of the duration of the flare, this corresponds to an energy dissipation rate in accelerated electrons of more than 100 times the stellar bolometric luminosity. The hard X-rays are also interpreted as lasting for timescales comparable with the flare soft X-ray emission, at variance with the Sun where this non-thermal component, when present, is generally impulsive. Regardless of the true nature of the hard X-ray flux, the equivalent width measured from the II~Peg {\\em Swift} spectra ranges between 18 and 60~eV, corresponding to efficiency values $\\mathcal{E}$ between 1\\% and 2\\%. These values are well within the range found from the theoretical calculations presented here, and comparable with the measured value for \\hr. Since the non-thermal hard X-ray component in the II~Peg flare is not well-constrained by the {\\em Swift} data and could also be explained by thermal emission, the observed Fe~K$\\alpha$ line seems more easily explained as arising from photoionization by flare X-rays (see also the discussion of D07b)." }, "0801/0801.3585_arXiv.txt": { "abstract": "We present a new non-parametric deprojection algorithm DOPING (Deprojection of Observed Photometry using and INverse Gambit), that is designed to extract the three dimensional luminosity density distribution $\\rho$, from the observed surface brightness profile of an astrophysical system such as a galaxy or a galaxy cluster, in a generalised geometry, while taking into account changes in the intrinsic shape of the system. The observable is the 2-D surface brightness distribution of the system. While the deprojection schemes presented hitherto have always worked within the limits of an assumed intrinsic geometry, in DOPING, geometry and inclination can be provided as inputs. The $\\rho$ that is most likely to project to the observed brightness data is sought; the maximisation of the likelihood is performed with the Metropolis algorithm. Unless the likelihood function is maximised, $\\rho$ is tweaked in shape and amplitude, while maintaining positivity, but otherwise the luminosity distribution is allowed to be completely free-form. Tests and applications of the algorithm are discussed. ", "introduction": "\\label{sec:intro} \\noindent The preliminary step involved in the dynamical modelling of galaxies, concerns the deprojection of the observed surface brightness distribution into the intrinsic luminosity density, as has been practised by [\\refcite{krajnovic04}], [\\refcite{kronawitter00}], [\\refcite{magog98}], among others. Deprojection though, is a non-unique problem, unless performed under very specific configurations of geometry and inclination, as discussed by [\\refcite{binneygerhard}], [\\refcite{kochanekrybicki}], [\\refcite{rybicki87}], [\\refcite{vandenbosch}] and others. Over the years, several deprojection schemes have been advanced and implemented within the purview of astronomy; these include parametric formalisms designed by [\\refcite{bendinelli}], [\\refcite{palmer}] and [\\refcite{cappellari}], as well as non-parametric methods, such as the Richardson-Lucy Inversion scheme developed by [\\refcite{richardson}] and [\\refcite{lucy}] and a method suggested by [\\refcite{romkoch}]. While the parametric schemes are essentially unsatisfactory owing to the dependence of the answer on the form of the parametrisation involved, the non-parametric schemes advanced till now have suffered from the lack of transparency and in the case of the Richardson-Lucy scheme, lack of an objective convergence criterion. Here, we present a new, robust non-parametric algorithm: Deprojection of Observed Photometry using an INverse Gambit (DOPING). DOPING does not need to assume axisymmetry but can work in a triaxial geometry with assumed axial ratios, and is able to incorporate radial variations in eccentricity. Although the code can account for changes in position angle, this facet has not been included in the version of the algorithm discussed here. Also, here we present the 1-D results obtained with DOPING but the code provides the full 3-D density distribution that projects to observed brightness map. In a future contribution, (Chakrabarty $\\&$ Ferrarese, {\\it in preparation}), DOPING will be applied to recover the intrinsic luminosity density of about 100 early type galaxies observed as part of the ACS Virgo Cluster Survey, as reported in [\\refcite{coteacs04}]. The paper has been arranged as follows. The basic framework of DOPING is introduced in Section~\\ref{sec:method}. This is followed by a short discourse on a test of the algorithm. An application to the observed data of the galaxy vcc1422 is touched upon in Section~\\ref{sec:1422}. Another application of DOPING is discussed in Section~\\ref{sec:clusters}. The paper is rounded up with a summary of the results. ", "conclusions": "\\noindent In this paper, we have introduced a new non-parametric algorithm DOPING that is capable of inverting observed surface brightness distributions of galaxies and galaxy clusters, while taking into account variations in the intrinsic shapes of these systems. The potency of DOPING is discussed in the context of a test galaxy in which the eccentricity is made to change radically with radius. The code is also successfully applied to obtain the luminosity density distribution of the faint nucleated galaxy vcc1422. Lastly, a novel use is made of the capability of DOPING to deproject in general geometries, in determining the intrinsic shape and inclination of a galaxy cluster. It is envisaged that implementing a measure of the LOS extent of a cluster from Sunyaev Zeldovich measurements, will help tighten the estimates of cluster inclination and the intrinsic axial ratios of triaxial clusters." }, "0801/0801.1255_arXiv.txt": { "abstract": "The nature of the extended hard X-ray source XMMU~J061804.3+222732 and its surroundings is investigated using {\\sl XMM-Newton}, {\\sl Chandra}, and {\\sl Spitzer} observations. This source is located in an interaction region of the IC~443 supernova remnant with a neighboring molecular cloud. The X-ray emission consists of a number of bright clumps embedded in an extended structured non-thermal X-ray nebula larger than 30\\arcsec in size. Some clumps show evidence for line emission at $\\sim$ 1.9 keV and $\\sim$ 3.7 keV at the 99\\% confidence level. Large-scale diffuse radio emission of IC~443 passes over the source region, with an enhancement near the source. An IR source of about 14\\arcsec $\\times$ 7\\arcsec size is prominent in the 24 $\\mu$m, 70 $\\mu$m, and 2.2 $\\mu$m bands, adjacent to a putative Si K-shell X-ray line emission region. The observed IR/X-ray morphology and spectra are consistent with those expected for J/C-type shocks of different velocities driven by fragmented supernova ejecta colliding with the dense medium of a molecular cloud. The IR emission of the source detected by {\\sl Spitzer} can be attributed to both continuum emission from an HII region created by the ejecta fragment and line emission excited by shocks. This source region in IC~443 may be an example of a rather numerous population of hard X-ray/IR sources created by supernova explosions in the dense environment of star-forming regions. Alternative Galactic and extragalactic interpretations of the observed source are also discussed. ", "introduction": "The energy release and the ejection of nucleosynthesis products by supernovae (SNe) events are of great importance for our understanding of the physics of the interstellar medium (ISM). The mixing of the ejected metals with the surrounding matter is of special interest when a SN occurs in a molecular cloud, which may cause further star-forming activity. Optical and UV studies of the structure of SN remnants (SNRs) have revealed a complex metal composition of ejecta and the presence of isolated high-velocity ejecta fragments interacting with surrounding media. The most prominent manifestations of this phenomena are the fast moving knots observed in some young ``oxygen-rich'' SNRs, such as the Galactic SNRs Cas A (e.g., Chevalier \\& Kirshner 1979; Fesen\\etal 2006), Puppis A (Winkler \\& Kirshner 1985), G292.0+1.8 (e.g. Winkler \\& Long 2006), and also N132D in the LMC and 1E 0102.2--7219 in the SMC (e.g. Blair et al. 2000). Ballistically moving ejecta fragments of SNRs can be considered as a class of hard X-ray sources. The prototype was observed in the Vela SNR (Aschenbach, Egger, \\& Tr\\\"umper 1995; Miyata et al.\\ 2001). A massive individual fragment moving supersonically through a molecular cloud can have a luminosity $L_x \\gsim$10$^{31} \\ergs$ in the 1--10 keV band, and is observable with \\xmm\\ and \\chan\\ at a few kpc distance (Bykov 2002, 2003). Its X-ray emission is expected to consist of two components. The first one is thermal X-ray emission from the hot shocked ambient gas behind the fragment bow shock, with a spectrum of an optically thin thermal plasma of an ISM-cloud abundance. The second emission component is nonthermal; the interaction of fast electrons accelerated at the fragment bow-shock with the fragment body produces a hard continuum as well as line emission (X-ray and IR), including the K-shell lines of Si, S, Ar, Ca, Fe, and other elements ejected by SN. Detection of the X-ray line emission would help distinguish an ejecta fragment from the other possible source of hard continuum emission associated with a SNR, namely, a pulsar wind nebula (PWN). A young SNR of an age of a few thousands years interacting with a molecular cloud can produce hundreds of X-ray sources associated with isolated ejecta fragments. They should be particularly numerous in starforming regions like those in the Galactic center region, where young core-collapsed supernovae in or near molecular clouds are expected to be present in abundance. The expected observational appearance of isolated ejecta fragments in a molecular cloud differs from what is seen in the Vela SNR. Ejecta fragments interacting with a dense molecular cloud are slowed down and crushed, and they are generally more bright. We will argue here that the X-ray emission spectra of ejecta fragments in a molecular cloud may be dominated by hard non-thermal components, because a powerful but very soft thermal component could be heavily absorbed. On the other hand, the spectra of fast supernova ejecta fragments propagating in a tenuous plasma, as it is the case in the Vela SNR, would be long-lived, less luminous and dominated by thermal emission. The present paper focuses on IC~443. This is a SNR of a medium age, estimated by Chevalier (1999) to be $\\sim$ 30,000 years, for which the number of X-ray sources from ejecta fragments should be {\\it much smaller\\ } than in a young SNR (possibly, only a few). It is, however, the best and most reliable laboratory to study this phenomenon since there are only very few examples of clearly established SNR-cloud interactions. IC~443 (G189.1+3.0) is an evolved SNR of about 45$\\arcmin$ size at a distance of 1.5 kpc (e.g. Fesen \\& Kirshner 1980). Radio observations of IC~443 (e.g. Braun \\& Strom 1986; Green 1986; Leahy 2004) show two half-shells. This appearance is probably due to interaction of the SNR with a molecular cloud that seems to separate the two half-shells. The molecular-cloud material has a torus-like structure (Cornett, Chin \\& Knapp 1977; Burton\\etal 1988; Troja, Bocchino, \\& Reale 2006), that can be interpreted as a sheet-like cloud first broken by the expanding pre-supernova wind and then by the SNR blast wave. Plenty of evidence for shock-excited molecules in this region has been found (e.g. DeNoyer 1979; Burton et al.\\ 1988; Dickman\\etal 1992; Turner\\etal 1992; van Dishoeck, Jansen, \\& Phillips 1993; Tauber\\etal 1994; Richter, Graham, \\& Wright 1995; Cesarsky\\etal 1999; Snell\\etal 2005). The complex structure of the interaction region, with evidence for multiple dense clumps, is seen in 2MASS images (e.g. Rho et al.\\ 2001). Three OH (1720 MHz) masers were found in IC~443 (Claussen \\etal 1997; Hewitt \\etal 2006, and references therein). Soft X-ray maps of IC~443 based on {\\sl ROSAT} data (Asaoka \\& Aschenbach 1994) and recent radio observations (Leahy 2004) suggest that another SNR, G189.6+3.3, is seen in the IC~443 field (see also the \\xmm\\ study by Troja, Bocchino, \\& Reale 2006). This makes the multiwavelength observational picture even more complex to interpret. The field of IC~443 was observed in X-rays with {\\sl HEAO 1\\/} (Petre et al.\\ 1988), {\\sl Ginga\\/} (Wang et al.\\ 1992), {\\sl ROSAT\\/} (Asaoka \\& Aschenbach 1994), {\\sl ASCA\\/} (Keohane et al.\\ 1997; Kawasaki et al.\\ 2002), {\\sl BeppoSAX\\/} (Preite-Martinez et al.\\ 2000; Bocchino \\& Bykov 2000), \\chan\\ (Olbert \\etal 2001; Bykov, Bocchino, \\& Pavlov 2005; Gaensler \\etal 2006; Weisskopf \\etal 2007), \\xmm\\ (Bocchino \\& Bykov 2001, 2003; Troja, Bocchino, \\& Reale 2006), and {\\sl RXTE} (Sturner, Keohane, \\& Reimer 2004). The X-ray emission of IC~443 below 4 keV is dominated by a number of thermal components (e.g., Petre \\etal 1988; Asaoka \\& Aschenbach 1994; Kawasaki et al.\\ 2002; Troja, Bocchino, \\& Reale 2006). The thermal-emission morphology is center-filled, with soft emission filaments visible at energies below 0.5 keV. A gradient of X-ray surface brightness at the SNR limb was found, as well as strong variations of absorbing column density \\nh, which indicates the complex molecular-cloud environment of IC~443 in the southern part of the remnant (e.g. Asaoka \\& Aschenbach 1994). {\\sl ASCA} observations have established that the hard X-ray emission of IC~443 (above 4 keV) is dominated by localized sources in the southern part of the remnant (Keohane et al.\\ 1997). In \\xmm\\ observations Bocchino \\& Bykov (2003; BB03 hereafter) found 12 sources with fluxes over 10$^{-14}~\\enf$ in the 2--10 keV band. Six of the detected sources are located in a relatively small, of $15^\\prime\\times 15^\\prime$ size, region projected onto the molecular cloud in the South-Eastern part of IC~443. {\\sl BeppoSAX\\/} MECS observations (4--10 keV) showed two sources, 1SAX~J0617.1+2221 and 1SAX~J0618.0+2227, with evidence from the {\\sl BeppoSAX\\/} PDS for the presence of hard emission up to 100 keV for the former (Bocchino \\& Bykov 2000). Observations of this source by \\chan\\ (Olbert\\etal 2001; Gaensler\\etal 2006; Weisskopf\\etal 2007) and \\xmm\\ (Bocchino \\& Bykov 2001) established its plerionic nature. Leahy (2004) argued that the pulsar that powers this plerion is associated with G189.6+3.3 rather than IC~443. The nature of the second hard source -- 1SAX~J0618.0+2227 -- remained unknown. This source, the brightest in the region (excluding the plerion), was resolved with \\xmm\\ into two sources -- the extended XMMU~J061804.3+222732 (of $\\sim$ 20\\arcsec size) and the point-like XMMU~J061806.4+222832. We will call them Src~1 and Src~2 respectively (note that the sources were listed as Src~11 and Src~12 in BB03). The position of XMMU~J061804.3+222732 in the remnant is illustrated in Figures~\\ref{xmm_spitzer} and \\ref{chan_xmm}. \\begin{figure*} \\includegraphics[width=0.99\\textwidth]{f1.eps} \\caption{ Wide-field views of SNR IC~443. {\\em Left:} {\\sl XMM-Newton} 2--8 keV image with {\\sl Spitzer} MIPS 24 $\\mu$m contours overlaid. {\\em Right:} {\\sl Spitzer} MIPS image at 24 $\\mu$m with {\\sl VLA} 1.4 GHz contours overlaid. The images are produced from the {\\sl XMM-Newton} observations 0114100101--0114100601 and 0301960101, and {\\sl Spitzer} MIPS observations r4616960, r4617216, and r4617472. The white arrow points to the studied region.} \\label{xmm_spitzer} \\end{figure*} A dedicated \\chan\\ observation of Src~1 has revealed a complex structure of a few bright clumps embedded in extended emission of $>$ 20\\arcsec size (Bykov, Bocchino, \\& Pavlov 2005; BBP05 hereafter). The brightest clumps are the extended Src~1a and the point-like Src~1b. The apparent position of the source in a SNR -- molecular cloud interaction region naturally leads to SNR-related interpretations. The observed X-ray morphology of Src\\,1 and the spectra of its components are consistent with expectations for a SN ejecta fragment interacting with a dense ambient medium. Alternatively, Src\\,1 could be interpreted as a PWN associated with either IC~443 or G189.6+3.3 (BBP05). However, one cannot exclude the extragalactic origin of the source, that is discussed in some detail in Section~\\ref{altern}. IC~443 is a candidate counterpart of the EGRET $\\gamma$-ray source 3EG~J0617+2238, with a flux of about 5$\\times$10$^{-7}$ cm$^{-2}$ s$^{-1}$ above 100 MeV (Esposito et al.\\ 1996). The spectrum of Src~1 extrapolated into the EGRET range is consistent with that of 3EG~J0617+2238 (BBP05). Also the position of Src~1 is consistent (albeit marginally) with that of 3EG~J0617+2238. Such a $\\gamma$-ray luminosity can be expected for both the fragment and PWN interpretations. The forthcoming {\\sl GLAST\\ } mission (e.g. Johnson 2006) will be able to provide an accurate position and spectrum of 3EG~J0617+2238, thus helping to solve the issue. Src~1 lies far away from the 99\\% error circle of the TeV-regime source recently reported by {\\sl MAGIC} (Albert et al. 2007) in the Western part of IC~443 field. The apparent position of TeV {\\sl MAGIC} source is close to the 1720 MHz OH maser detected by Claussen et al. (1997). We present here new results of a deep 80 ks observation of the region with \\xmm , imaging of the region with the {\\sl Spitzer} infrared observatory, and a new analysis of VLA radio observations. In \\S\\,2 a combined analysis of the new \\xmm\\ observations and all the previous high-resolution X-ray data from \\chan\\ and \\xmm\\ is presented, including images, spectra, and time variations in the X-ray domain. In \\S\\,3 archival radio (VLA), IR (2MASS and {\\sl Spitzer} MIPS), and optical (POSS-II) data are used to constrain the nature of Src~1. A discussion of the obtained results and future prospects are presented in \\S\\,4. ", "conclusions": "\\label{concl} The multi-wavelength observations presented here indicate a possible physical connection of the X-ray source J0618 with the neighboring IR {\\sl Spitzer} and 2MASS sources. That connection, if real, can be understood in a scenario where J0618 originates in an interaction of the IC~443 SNR with the adjacent molecular cloud. The correlation would require the presence of both fast and slow shocks in the clumpy molecular-cloud material. The X-ray line features apparent in the spectra of the clumps favor a scenario in which the shocks are produced by a fast ballistically moving SN ejecta fragment penetrating into a structured molecular cloud. The model provides a physical picture coherent with the current observational data, although alternative scenario cannot be rejected yet. Alternatively, Src 1 can be interpreted as a massive X-ray cluster of galaxies at a redshift $z >$ 0.5. High-resolution arcsecond-scale observations and fine spectroscopy are required to distinguish between these very different scenarios. If the SNR-ejecta interpretation is confirmed by further observations, the source J0618 in the IC~443 could be a prototype of a rather numerous population of hard X-ray -- IR sources created by SN explosions in the dense environments of star-forming regions. Such sources would be particularly abundant in the Galactic Centre region." }, "0801/0801.3250_arXiv.txt": { "abstract": "The Alpha Magnetic Spectrometer (AMS), whose final version AMS-02 is to be installed on the International Space Station (ISS) for at least 3 years, is a detector designed to measure charged cosmic ray spectra with energies up to the TeV region and with high energy photon detection capability up to a few hundred GeV. It is equipped with several subsystems, one of which is a proximity focusing RICH detector with a dual radiator (aerogel+NaF) that provides reliable measurements for particle velocity and charge. The assembly and testing of the AMS RICH % is currently being finished % and the full AMS detector is expected to be ready by the end of 2008. The RICH detector of AMS-02 is presented. Physics prospects are briefly discussed. ", "introduction": "The Alpha Magnetic Spectrometer (AMS)\\cite{bib:ams}, whose final version AMS-02 is to be installed on the International Space Station (ISS) for at least 3 years, is a detector designed to study the cosmic ray flux by direct detection of particles above the Earth's atmosphere, at an altitude of \\mbox{$\\sim$ 400 km}, using state-of-the-art particle identification techniques. AMS-02 is equipped with a superconducting magnet cooled by superfluid helium. The spectrometer is composed of several subdetectors: a Transition Radiation Detector (TRD), a Time-of-Flight (ToF) detector, a Silicon Tracker, Anticoincidence Counters (ACC), a Ring Imaging \\CK\\ (RICH) detector and an Electromagnetic Calorimeter (ECAL). A preliminary version of the detector, AMS-01, was successfully flown aboard the US space shuttle Discovery in June 1998. \\begin{figure}[htb]% \\center \\vspace{-0.5cm} \\begin{tabular}{cc} \\mbox{\\epsfig{file=amsrich.eps,width=0.47\\textwidth,clip=}} & \\mbox{\\epsfig{file=RICHmaravilhoso_new.ps,width=0.43\\textwidth,clip=, bbllx=-53,bblly=100,bburx=667,bbury=666}} \\end{tabular} \\vspace{-0.2cm} \\caption{The RICH detector of AMS-02 \\emph{(left)}. View of the assembled RICH detector at CIEMAT \\emph{(right)}.\\label{richdet}} \\vspace{-0.2cm} \\end{figure} The main goals of the AMS-02 experiment are: (i) a precise measurement of charged cosmic-ray spectra in the rigidity region between \\mbox{$\\sim$ 0.5 GV} and \\mbox{$\\sim$ 2 TV} and the detection of photons with energies up to a few hundred GeV; (ii) a search for heavy antinuclei \\mbox{($Z \\ge$ 2)}, which if discovered would signal the existence of cosmological antimatter; (iii) a search for dark matter constituents by examining possible signatures of their presence in cosmic ray spectra. The long exposure time and large acceptance \\mbox{(0.5 m${}^2\\cdot$sr)} of AMS-02 will enable it to collect an unprecedented statistics of more than $10^{10}$ nuclei. \\vspace{-0.3cm} ", "conclusions": "AMS-02 will provide a new insight on the cosmic-ray spectrum by collecting precise data for an unprecedented number of particles above the Earth's atmosphere. The RICH detector will play a key role in the operation of AMS due to its capabilities for velocity reconstruction, charge determination and albedo rejection. Extensive testing has been performed on the RICH detector and its components. The assembly of the RICH detector is currently being finished and the full AMS detector is expected to be ready by the end of 2008. \\vspace{-0.4cm}" }, "0801/0801.0645_arXiv.txt": { "abstract": "Observations from the Space Telescope Imaging Spectrograph define the flux of the DBQ4 star LDS749B from 0.12--1.0~$\\mu$m with an uncertainty of $\\sim$1\\% relative to the three pure hydrogen WD primary \\emph{HST} standards. With $T_\\mathrm{eff}=13575~K$, $\\log g=8.05$, and a trace of carbon at $<$1$\\times10^{-6}$ of solar, a He model atmosphere fits the measured STIS fluxes within the observational noise, except in a few spectral lines with uncertain physics of the line broadening theory. Upper limit to the atmospheric hydrogen and oxygen fractions by number are 1$\\times10^{-7}$ and 7$\\times10^{-10}$, respectively. The excellent agreement of the model flux distribution with the observations lends confidence to the accuracy of the modeled IR fluxes beyond the limits of the STIS spectrophotometry. The estimated precision of $\\sim$1\\% in the predicted IR absolute fluxes at 30~$\\mu$m should be better than the model predictions for Vega and should be comparable to the absolute accuracy of the three primary WD models. ", "introduction": "The DBQ4 star LDS749B (WD2129+00) has long been considered for a flux standard (e.g., Bohlin et~al.\\ 1990). To establish the flux on the \\emph{Hubble Space Telescope} (\\emph{HST}) white dwarf (WD) flux scale, STIS spectrophotometry was obtained in 2001--2002. The virtues of LDS749B as a flux standard include an equatorial declination and a significantly cooler flux distribution than the 33000--61000~K primary DA standards GD71, GD153, and G191B2B. Full STIS wavelength coverage is provided from 0.115--1.02~$\\mu$m, and the peak in the SED is near 1900~\\AA. At $V=14.674$ (Landolt \\& Uomoto 2007), LDS749B is among the faintest \\emph{HST} standards and is suitable for use with larger ground-based telescopes and with the more sensitive \\emph{HST} instrumentation, such as the ACS/SBC and COS. The bulk of the STIS data was obtained as part of the FASTEX (Faint Astronomical Sources Extention) program. Finding charts appear in Turnshek et~al.\\ (1990) and in Landolt \\& Uomoto (2007); but there is a large proper motion of 0.416 and 0.034 arcsec/yr in right ascension and declination, respectively. The absolute flux calibration of \\emph{HST} instrumentation is based on models of three pure hydrogen WD stars GD71, GD153, and G191B2B (Bohlin 2000; Bohlin, Dickenson, \\& Calzetti 2001; Bohlin 2003). In particular, the NLTE model fluxes produced by by the Tlusty code (Hubeny \\& Lanz 1995) determine the shape of the flux distributions using the known physics of the hydrogen atom and of stellar atmospheres. If there are no errors in the basic physics used to determine the stellar temperatures and gravities from the Balmer line profiles, then the uncertainty of 3000~K for the effective temperature of G191B2B means that the relative flux should be correct to better than 2.5\\% from 0.13 to 1~$\\mu$m and to better than 1\\% from 0.35 to 1~$\\mu$m. A model that matches the observations serves as a noise free surrogate for the observational flux distribution and provides a reliable extrapolation beyond the limits of the observations for use as a calibration standard for \\emph{JWST}, \\emph{Spitzer}, and other IR instrumentation. Currently, the best IR absolute flux distributions are found in a series of papers from the epic and pioneering work of M.~Cohen and collaborators, i.e., the Cohen-Walker-Witteborn (CWW) network of absolute flux standard (e.g., Cohen, Wheaton, \\& Megeath 2003; Cohen 2007). The CWW IR standard star fluxes are all ultimately based on models for Vega and Sirius (Cohen et~al.\\ 1992). More recently, Bohlin \\& Gilliland (2004) observed Vega and published fluxes on the \\emph{HST}/STIS WD flux scale. A small revision in the STIS calibration resulted in excellent agreement of the STIS flux distribution with a custom made Kurucz model with $T_\\mathrm{eff}=9400$~K (Bohlin 2007), which is the same $T_\\mathrm{eff}$ used for the Cohen et~al.\\ (1992) Vega model. The model presented here for LDS749B and archived in the CALSPEC database\\footnote{The absolute spectral energy distributions discussed in this paper are available in digital form at http://www.stsci.edu/hst/observatory/cdbs/calspec.html.} should have a better precision than the Kuruzc $T_\\mathrm{eff}=9400$~K model for Vega, especially beyond $\\sim$12~$\\mu$m, where the Vega's dust disk becomes important (Engleke, Price, \\& Kraemer 2006). Vega is also a pole-on rapid rotator, which may also cause IR deviations from the flux for a single temperature model. Our modeled flux distribution for LDS749B should have an accuracy comparable to the pure hydrogen model flux distributions for the primary WD standards GD71, GD153, and G191B2B. ", "conclusions": "In the absence of any interstellar reddening, a helium model with $T_\\mathrm{eff}=13575~K\\pm50$, $\\log g=8.05\\pm0.7$, and a trace of carbon at $<$1$\\times10^{-6}$ of solar fits the measured STIS flux distribution for LDS749B. The noise-free, absolute flux distribution from the model after normalization to the observed broadband visual flux is preferred for most purposes. This normalized model SED is a high fidelity far-UV to far-IR calibration source; and the flux distribution is available via Table~3 in the electronic version of the \\emph{Journal}. Both the observed flux distribution and the modeled fluxes are also available from the CALSPEC database.\\footnote{http://www.stsci.edu/hst/observatory/cdbs/calspec.html/.}" }, "0801/0801.2476_arXiv.txt": { "abstract": "{ The chemical properties of galaxies and their evolution as a function of cosmic epoch are powerful constraints of their evolutionary histories. } { This work provides a grid of numerical models of galaxy evolution over an extended cosmic epoch. The aims are to assess how well current models reproduce observed properties of galaxies, in particular the stellar mass versus gas phase metallicity relation, and to quantify the effect of the merging histories of galaxies on their final properties. } { We use 112 N-body/hydrodynamical simulations in the standard Cold Dark Matter universe, to follow the formation of galaxy-sized halos and investigate the chemical enrichment of both the stellar component and the interstellar medium of galaxies, with stellar masses larger than $\\sim10^9$~M$_{\\odot}$. } { The resulting chemical properties of the simulated galaxies are broadly consistent with the observations. The predicted relationship between the mean metallicity and the galaxy stellar mass for both the stellar and the gaseous components at $z=0$ are in agreement with the relationships observed locally. The predicted scatter about these relationships, which is traced to the differing merging histories amongst the simulated galaxies with similar final masses, is similar to that observed. Under the hierarchical formation scenario, we find that the more massive galaxies are typically more evolved than their low mass counterparts over the second half of the age of the Universe. The predicted correlations between the total mass and the stellar mass of galaxies in our simulated sample from the present epoch up to $z\\sim 1$ agree with observed ones. We find that the integrated stellar populations in the simulations are dominated by stars as old as $4-10$~Gyr. In contrast with massive galaxies, for which the luminosity-weighted ages of the integrated stellar populations in the simulated sample agree with those derived from the modeling of observed spectral energy distributions, simulated galaxies with stellar masses $\\sim10^{9}$~M$_{\\odot}$ at $z=0$ tend to be older than the local galaxies with similar stellar masses. } { The stellar mass versus metallicity relation and its associated scatter are reproduced by the simulations as consequences of the increasing efficiency of the conversion of gas into stars with stellar mass, and the differing merging histories amongst the galaxies with similar masses. The old ages of simulated low mass galaxies at $z=0$, and the weak level of chemical evolution for massive galaxies suggest however that our modelling of the supernova feedback may be incomplete, or that other feedback processes have been neglected. } ", "introduction": "The chemical enrichment histories of galaxies provide insight into various processes involved in galaxy formation and evolution. The chemical composition of stars and gas within a galaxy depends on a number of physical processes, such as the star formation history, gas outflows and inflows, stellar initial mass function, etc. Although it is a complicated task to disentangle the effects of these processes, the galactic chemical abundances at various epochs place tight constraints on the likely evolutionary histories of galaxies. The correlation between galaxy metallicity and luminosity in the local universe is one of the most significant observational results in galaxy chemical evolution studies. Lequeux et al. (1979) first revealed that the oxygen abundance increases with the total mass of irregular galaxies. The luminosity-metallicity relation for irregulars was later confirmed by Skillman et al. (1989) and Richer \\& McCall (1995), amongst others. Subsequent studies have extended the relation to spiral galaxies (e.g., Garnett \\& Shields 1987; Zaritsky et al. 1994; Garnett et al. 1997; Pilyugin \\& Ferrini 2000), and to elliptical galaxies (Brodie \\& Huchra 1991). The luminosity correlates with metallicity over $\\sim10$ magnitudes in luminosity and a factor of $\\sim100$ in metallicity, with indications suggesting that the relationship may be independent of environment (Vilchez 1995; Mouhcine et al. 2007) and morphology (Mateo 1998). More recently, large samples of star-forming galaxies drawn from galaxy redshift surveys, e.g. 2dF Galaxy Redshift Survey and the Sloan Digital Sky Survey (SDSS hereafter), have been used to confirm the existence of the luminosity-metallicity relation over a broad range of luminosity and metallicity (Lamareille et al. 2004; Tremonti et al. 2004). Lee et al. (2006) have extended the mass-metallicity relation to dwarf irregular galaxies, and found that the dispersion is similar over five orders of magnitude in stellar mass, and that the relation between the integrated stellar mass and the oxygen abundance of the interstellar medium is similarly tight from high stellar mass to low stellar mass galaxies. {\\bf The gas phase oxygen abundance vs. stellar mass has been understood either as a depletion sequence or a sequence in astration (see e.g. Tremonti et al. 2004). K\\\"oppen, Weidner, \\& Kroupa (2007) have presented an alternative explanation of the mass-metallicity relation arguing it could be the consequence of the integrated galactic initial mass function depending on the star formation rate, leading to higher oxygen yield in systems where the star formation rate is high. More massive galaxies are expected to have higher yields and thus have a tendency to have higher metallicities. Dalcanton (2007) presented a series of closed-box chemical evolution models including infall and outflow. She has shown that neither simple infall nor outflow can reproduce the observed low effective yields in low-mass galaxies (see Lee et al. 2006 for a similar conclusion), but metal-enriched outflows can do. That is, effectively the freshly synthesized elements need to be removed from the matter cycle, which is, as noted by K\\\"oppen et al. (2007) in principle, equivalent to reducing the number of massive stars.} Many recent studies in galaxy evolution trace changes in scaling relations of galaxies at earlier epochs. In this context, the galactic chemical abundances at different cosmic epochs can assist in constraining the likely scenarios of galaxy evolution. Different groups have used the classical nebular diagnostic techniques developed to study the properties of {H{\\sc ii}} regions and emission line galaxies in the local universe to probe the properties of the interstellar gas in intermediate ($010^{10}$ M$_{\\sun}$) galaxies. Lower mass galaxies have greater scatter. For some simulated galaxies, star formation is heavily suppressed and leads to lower stellar mass compared with the total mass. \\item[-] We have analyzed the metallicity for both stellar and gas components. Both are correlated with the stellar mass up to $z\\sim 1.2$, and show an agreement with recent observations (Figs.~\\ref{oh_ms_gas} and \\ref{zs_mstar}). This relation arises from the fact that higher mass galaxies convert gas more efficiently into stars. \\item[-] Higher mass galaxies reach a high metallicity earlier. As a result, the slope of the mass-metallicity relation becomes flatter with decreasing redshift (Figs.~\\ref{oh_ms_gas} and \\ref{zs_mstar}). Lower mass galaxies (${\\rm M_{tot}}<10^{10}$ M$_{\\sun}$) evolve chemically more slowly, and have lower metallicity than that predicted from the extrapolation of the mass-metallicity relation for higher mass galaxies at lower redshifts, which is also consistent with the observational data. \\item[-] The predicted scatter in the mass-metallicity relation also reproduces qualitatively the observed scatter (Figs.~\\ref{oh_ms_gas} and \\ref{zs_mstar}). This suggests that the observed scatter could be due to the difference in mass assembly histories, a natural prediction of the hierarchical clustering scenario (Fig.~\\ref{mass_assembly}). \\end{itemize} The primary disagreement between our simulations and observations is as follows: \\begin{itemize} \\item[-] Our simulations predict almost no correlation between the stellar age and the stellar mass, which cannot explain the observed significantly younger population in lower mass galaxies, e.g. stellar masses lower than ${\\rm \\sim 10^{9} M_{\\odot}}$ (Fig.~\\ref{agestar_mstar}). As a result, the simulations also predict too flat of a colour-luminosity relation (Fig.~\\ref{cmr}). \\end{itemize} This suggests that the simulations need a mechanism to sustain a low level of star formation activity and also keep the stellar and gas metallicity low. A combination of stronger supernovae feedback and the UV background radiation may help to cause such self-regulated star formation in the lower mass galaxies. This demonstrates that the quantitative comparisons made in the paper are invaluable to improve numerical simulation models, which eventually will aid in completing our understanding of the physical processes governing galaxy formation and evolution." }, "0801/0801.0968_arXiv.txt": { "abstract": "Clusters of galaxies are self-gravitating systems of mass $\\sim 10^{14}-10^{15} h^{-1}$ M$_\\odot$ and size $\\sim 1-3 h^{-1}$ Mpc. Their mass budget consists of dark matter ($\\sim 80\\%$, on average), hot diffuse intracluster plasma ($\\lesssim 20\\%$) and a small fraction of stars, dust, and cold gas, mostly locked in galaxies. In most clusters, scaling relations between their properties, like mass, galaxy velocity dispersion, X-ray luminosity and temperature, testify that the cluster components are in approximate dynamical equilibrium within the cluster gravitational potential well. However, spatially inhomogeneous thermal and non-thermal emission of the intracluster medium (ICM), observed in some clusters in the X-ray and radio bands, and the kinematic and morphological segregation of galaxies are a signature of non-gravitational processes, ongoing cluster merging and interactions. Both the fraction of clusters with these features, and the correlation between the dynamical and morphological properties of irregular clusters and the surrounding large-scale structure increase with redshift. In the current bottom-up scenario for the formation of cosmic structure, where tiny fluctuations of the otherwise homogeneous primordial density field are amplified by gravity, clusters are the most massive nodes of the filamentary large-scale structure of the cosmic web and form by anisotropic and episodic accretion of mass, in agreement with most of the observational evidence. In this model of the universe dominated by cold dark matter, at the present time most baryons are expected to be in a diffuse component rather than in stars and galaxies; moreover, $\\sim 50\\%$ of this diffuse component has temperature $\\sim 0.01-1$ keV and permeates the filamentary distribution of the dark matter. The temperature of this Warm-Hot Intergalactic Medium (WHIM) increases with the local density and its search in the outer regions of clusters and lower density regions has been the quest of much recent observational effort. Over the last thirty years, an impressive coherent picture of the formation and evolution of cosmic structures has emerged from the intense interplay between observations, theory and numerical experiments. Future efforts will continue to test whether this picture keeps being valid, needs corrections or suffers dramatic failures in its predictive power. ", "introduction": "The present chapter provides the general framework of the physics of clusters of galaxies: it outlines how clusters connect to several astrophysical issues, from cosmology to the formation of galaxies and stars. Some of these topics are discussed in the chapters of this volume; we refer to these chapters when appropriate. When the topic we mention here is not dealt with elsewhere in this volume, we refer to some of the most recent reviews. ", "conclusions": "" }, "0801/0801.0703_arXiv.txt": { "abstract": "{} {We study the visibility of sunspots and its influence on observed values of sunspot region parameters.} {We use Virtual Observatory tools provided by AstroGrid to analyse a sample of 6862 sunspot regions. By studying the distributions of locations where sunspots were first and last observed on the solar disk, we derive the visibility function of sunspots, the rate of magnetic flux emergence and the ratio between the durations of growth and decay phases of solar active regions.} {We demonstrate that the visibility of small sunspots has a strong centre-to-limb variation, far larger than would be expected from geometrical (projection) effects. This results in a large number of young spots being invisible: 44\\% of new regions emerging in the west of the Sun go undetected. For sunspot regions that are detected, large differences exist between actual locations and times of flux emergence, and the apparent ones derived from sunspot data. The duration of the growth phase of solar regions has been, up to now, underestimated. } {} ", "introduction": "The birth of a new spot on the solar disk indicates the emergence of magnetic flux through the photosphere, a process which is key to the solar cycle \\cite{Sol2003,Fis2000} and the study of stellar magnetic dynamos. Sunspots also cause variations in the total solar irradiance, an important parameter in determining the Sun's influence on climate \\cite{Fou2006}. The presence of a sunspot is key to a solar region being assigned an active region number by the NOAA Space Weather Prediction Center (http://www.swpc.noaa.gov) so that its evolution and activity can be tracked \\cite{Gal2007}. The formation and evolution of active regions are fundamental to solar dynamic phenomena such as flares and Coronal Mass Ejections and their effect on the Earth environment. The visibility of sunspots is currently thought to be limited only by geometrical effects arising from projection of the solar sphere onto a 2D image, effects referred to as foreshortening. In this paper we present results obtained serendipitously while analysing sunspot data by means of Virtual Observatory tools, showing that the visibility of small sunspots is much poorer than predicted by the foreshortening model. ", "conclusions": "Figure 3-b demonstrates that the visibility of small sunspots is much poorer than expected from geometrical effects associated with foreshortening. It shows that the minimum area required for a spot to be detected at $\\lambda$=$\\pm$30$^{\\circ}$ is more than twice the threshold area at $\\lambda$=0$^{\\circ}$ if $k$=14 msh/day and almost 4 times if $k$=30 msh/day. The centre-to-limb variation in visibility is remarkably large. We investigated whether the asymmetries shown in Fig. 1 display any solar cycle dependence, and found no evidence of it. The USAF/Mount Wilson locations and times of first appearances agree with those from SOHO/MDI continuum data (as verified manually for a sample of regions). It is known that seeing associated with ground-based data does not cause significant reduction of visibility \\cite{Gyo2004}. How many spots are affected by the visibility effects here described? The distribution of sunspot areas measured at a single longitudinal location on the solar disk is lognormal \\cite{Bog1988} and the number of spots with area at Central Meridian around 10 msh is more than 2 orders of magnitude larger than the number of spots with area of about 100 msh. Therefore the majority of sunspots will cross the visibility curve shown in Fig. 3-b. Our results have a number of important implications. The first is that the radiative processes that make a small region of strong magnetic field appear as a dark spot, have a strong centre-to-limb variation. This may prove important for the study of sunspots' 3D structure and will require further investigation. Whether larger spots are also affected by the same process will also need further study. Reports of centre-to-limb variations of corrected sunspot areas have appeared in the literature \\cite{Gyo2004, Hoy1983}. Faculae have a large centre-to-limb variation, and their contrast changes sign as one moves towards the disk centre, resulting in their being darker than the surrounding photosphere at the disk centre \\cite{Law1993}. This demonstrates that the appearance of photospheric magnetic flux tubes strongly depends on the viewing angle. The second implication is that actual distributions of sunspot lifetimes and areas may differ from the apparent ones derived from observations. The latter have been used to constrain mechanisms of sunspot formation and decay \\cite{Sol2003, Pet1997, Mar1993}. A large number of regions reported of short duration may in fact have longer lifetimes, and be crossing in and out of the visibility curve. The Gnevyshev-Waldmeier law, stating that a sunspot's lifetime increases linearly with its maximum size, may need to be reassessed in light of our results. Poor visibility of region emergence means that the actual time of magnetic flux emergence can be much earlier than the apparent time, e.g.~for a region seen to emerge at $\\lambda$=$-$50$^{\\circ}$, by approximately 2 days. On the other hand, a large fraction of new emergences in the western portion of the solar disk go undetected. By using the data in the inset of Fig.~1-a for $\\lambda$$>$0 and the value of actual number of emergences $N_1$=160 obtained from the data of Fig. 2, we obtain that 44\\% of new spots emerging in [0$^{\\circ}$, $+$60$^{\\circ}$] were invisible. The presence of a sunspot is key to a solar active region being assigned an Active Region Number by the NOAA Space Weather Prediction Center (see e.g. Dalla et al. (2007) for the full list of criteria). A region that has been given a NOAA number is monitored and its activity tracked \\cite{Gal2007}. New regions emerging in the west of the Sun with their spots being invisible are missed and not tracked. We conclude that current criteria for assigning Active Region Numbers may need to be revised and that EUV solar images may need to be routinely used to supplement white-light information. Sunspots are well known to cause depletions in the total solar irradiance (TSI) \\cite{Fou2006} and many models of TSI variation need as input information on the number and areas of sunspots. % While the sunspots that most affect TSI are the largest ones, of area typically well above the visibility threshold shown in Fig. 3-b, our results impact TSI studies because they demonstrate that the apparent age and stage of development of a sunspot may not correspond to the actual ones. The latter information is required when studying the time dependence of sunspot effects on TSI and whether the age of a region is an important factor in determining the magnitude of TSI decrease. The asymmetry in the distribution of emergences was obtained, unexpectedly, during a study aiming at cross correlating catalogues of sunspot regions and flares, by means of AstroGrid workflows \\cite{Dal2007}. This demonstrates the usefulness of VO tools in making new science possible, by provision of better tools for analysis of large datasets." }, "0801/0801.1828_arXiv.txt": { "abstract": "Two recent issues realted to nucleosynthesis in early proton-rich neutrino winds are investigated. In the first part we investigate the effect of nuclear physics uncertainties on the synthesis of $^{92}$Mo and $^{94}$Mo. Based on recent experimental results, we find that the proton rich winds of the model investigated here can not be the only source of the solar abundance of $^{92}$Mo and $^{94}$Mo. In the second part we investigate the nucleosynthesis from neutron rich bubbles and show that they do not contribute to the nucleosynthesis integrated over both neutron and proton-rich bubbles and proton-rich winds. ", "introduction": "Over the past decade improvements in neutrino-transport and multi-dimensional computer simulations have lead to a new understanding of the conditions that lead to nucleosynthesis of the elements above iron in core-collapse supernovae. Immediately following the bounce on the proto-neutron star, the shock fully photodisintegrates the infalling material turning it into electron--position pairs, neutrons, and protons. As the nascent neutron star continues to collapse it liberates $10^{53} \\textrm{ergs}$ over the span of $\\sim 10$ seconds primarily in the form of neutrinos. This enormous neutrino flux is deposited in the low density region of photodisintegrated matter inside the gain radius between the neutron star and the accretion shock of the still infalling material and heats it to temperatures in excess of 10 billion K while driving mass away in the form of a neutrino wind theoretically leading to the explosion of the supernova \\cite{Qian96}. The strong flux of neutrinos and anti-neutrinos results in a detailed balance between protons and neutrons that favors the lighter mass protons depending on the respective neutrino spectra leading to an electron fraction that is proton-rich ($Y_e>0.5$) \\citep{Liebendoerfer03, Froehlich06b}. These protons and neutrons recombine into alpha particles that proceed via the $\\alpha(\\alpha n,\\gamma)$ ${}^{9}\\textrm{Be}(\\alpha,n)$ ${}^{12}\\textrm{C}$-reactions followed by a series of $(\\alpha,\\gamma)$-reactions or combined $(\\alpha,p)(p,\\gamma)$-reactions along $N=Z$ into the iron group, primarily ${}^{56}\\textrm{Ni}$ and ${}^{60}\\textrm{Zn}$ which form the seeds of the subsequent nucleosynthesis. From this point the resulting nucleosynthesis in the neutrino-driven wind essentially depends on the number of seed nuclei to the number of excess neutrons or protons that were frozen out and did not turn into seed nucleii ($Y_e$), the entropy per baryon, the expansion timescale of the ejecta and the amount of the ejecta. As the explosion evolves, an ejected mass element inherits some combination of these parameters and below $\\sim 0.5\\textrm{MeV}$ they remain fairly constant as the matter proceeds to freeze out. In this paper, we consider the early times when the wind still contains a proton excess because the rates for neutrino and positron captures on neutrons are faster than those for the inverse captures on protons. We consider two interesting problems which are discussed in the following two sections. ", "conclusions": "" }, "0801/0801.3745.txt": { "abstract": "We describe the HST ACS Coma cluster Treasury survey, a deep two-passband imaging survey of one of the nearest rich clusters of galaxies, the Coma cluster (Abell 1656). The survey was designed to cover an area of 740 arcmin$^2$ in regions of different density of both galaxies and intergalactic medium within the cluster. The ACS failure of January 27th 2007 leaves the survey 28\\% complete, with 21 ACS pointings (230 arcmin$^2$) complete, and partial data for a further 4 pointings (44 arcmin$^2$). Predicted survey depth for 10$\\sigma$ detections for optimal photometry of point sources is g$^{\\prime}$ = 27.6 in the F475W filter, and I$_C$=26.8 mag in F814 (AB magnitudes). Initial simulations with artificially injected point sources show 90\\% recovered at magnitude limits of g$^{\\prime}$ = 27.55 and I$_C$ = 26.65. For extended sources, the predicted 10$\\sigma$ limits for a 1 arcsecond$^{2}$ region are g$^{\\prime}$ = 25.8 mag arcsec$^{-2}$ and I$_C$ = 25.0 mag arcsec$^{-2}$. We highlight several motivating science goals of the survey, including study of the faint end of the cluster galaxy luminosity function, structural parameters of dwarf galaxies, stellar populations and their effect on colors and color gradients, evolution of morphological components in a dense environment, the nature of ultra compact dwarf galaxies, and globular cluster populations of cluster galaxies of a range of luminosities and types. This survey will also provide a local rich cluster benchmark for various well known {\\it global} scaling relations and explore new relations pertaining to the {\\it nuclear} properties of galaxies. ", "introduction": "\\label{sec:introduction} The Coma cluster of galaxies, Abell 1656, is along with the Perseus cluster the nearest rich, and dense cluster environment. Unlike the Perseus cluster, it is at high galactic latitude (b = 87.9$^{\\circ}$) and it has been a popular target for study at all wavelengths. Progressively deeper wide-area photometric surveys of Coma have become available over the past 30 years, and waveband coverage has spread from the original B and V band surveys into the near ultra-violet and infra-red (Godwin \\& Peach 1977; Godwin, Metcalfe \\& Peach 1983 (GMP); Komiyama et al.\\ 2002; Adami et al.\\ 2006; Eisenhardt et al.\\ 2007). A larger area around Coma is covered by the imaging part of Data Release 5 of the Sloan Digital Sky Survey (Adelman-McCarthy et al. 2007). From these surveys, samples of galaxies have been selected for spectroscopic study, which has resulted in an understanding of the internal dynamics of the cluster (Colless \\& Dunn 1996; Mobasher et al.\\ 2001; Edwards et al.\\ 2002; Gutierrez et al.\\ 2004), the internal dynamics of cluster members (Lucey et al. 1991; Jorgensen et al.\\ 1996; Smith et al.\\ 2004; Matkovic \\& Guzm\\'an 2005; Cody et al. 2007), and their mean luminosity weighted stellar ages, abundances and $\\alpha$-enhancement (Bower et al. 1992a,b; Guzm\\'an et al. 1992; Caldwell et al. 1993; Jorgensen 1999; Poggianti et al. 2001; Moore et al. 2002; Nelan et al. 2005; Smith et al. 2006; Matkovic et al. 2007). Coma presents us with the best opportunity to study large samples of galaxies of different luminosity, environment and morphological type, but at a common distance, and with a common Galactic extinction. There is good agreement on the distance to the Coma cluster, with independent studies using six different methods yielding values in the range 84 -- 108 Mpc, as summarised in Table \\ref{tab:Distances}. These values fit well with the current concordance cosmology: assuming $H_0$ = 71 km/s/Mpc; $\\Omega_{\\lambda}$ = 0.73; $\\Omega_m$ = 0.27, and a redshift z = 0.0231 gives a distance of 99.3 Mpc. In this paper we assume a distance of 100 Mpc, equivalent to a distance modulus of 35.00. \\clearpage \\begin{deluxetable}{lccl} \\tablecolumns{4} \\tablewidth{0pc} \\tablecaption{\\label{tab:Distances}The Distance to Coma determined by different techniques} \\tablehead{ \\colhead{Technique}&\\colhead{Distance (Mpc)}&\\colhead{Distance Modulus}&\\colhead{Reference}} \\startdata I-band Tully-Fisher&86.3$\\pm$6&34.68$\\pm$0.15&Tully \\& Pierce (2000)\\\\ K$^{\\prime}$-band SBF&85$\\pm$10&34.64$\\pm$0.27&Jensen et al. (1999)\\\\ I-band SBF&102$\\pm$14&35.04$\\pm$0.32&Thomsen et al. (1997)\\\\ $D_n-\\sigma$&96$\\pm$6&34.90$\\pm$0.14&Gregg (1997)\\\\ Fundamental Plane&108$\\pm$12&35.16$\\pm$0.25&Hjorth \\& Tanvir (1997)\\\\ Globular Cluster LF&102$\\pm$6&35.05$\\pm$0.12&Kavelaars et al. (2000)\\\\ \\enddata \\end{deluxetable} \\clearpage Coma lies in a rich region of the large-scale distribution of galaxies, at the intersection of a number of filaments. Figure \\ref{fig:coma_xyz} shows two projections of the distribution of galaxies in supergalactic co-ordinates in the region of Coma and the nearby richness class 2 cluster Abell 1367. The Great Wall (Geller \\& Huchra 1989) a vertical structure in the two panels of Figure \\ref{fig:coma_xyz}, runs through these two clusters, other filaments intersect the Great Wall at the Coma cluster. \\clearpage \\begin{figure*} \\begin{center} \\includegraphics[width=12cm]{f1.ps} \\caption{\\label{fig:coma_xyz}The location and environment of the Coma cluster in supergalactic co-ordinates. The left panel shows a projection onto supergalactic Y, close to the plane of the sky. The right panel shows a projection onto supergalactic X, in this panel the horizontal (Y) axis is close to measured redshift. All axes are in equivalent cz. Positions of the points are derived from measured sky positions and redshifts. Members of Coma (centre) and Abell 1367 (bottom) are plotted as filled red circles, for these galaxies the redshift used to compute the distance is the cluster redshift, with a random velocity offset chosen to make the cluster appear round. For non cluster members, measured cz is used, and the positions are plotted as filled black circles if they have velocities within $\\pm$600 km/s of the Coma mean in the depth dimension in that particular panel, and as open grey circles otherwise. The Great Wall is the structure at SGY $\\sim 6800$ km/s in the right panel and is seen face on in the left panel} \\end{center} \\end{figure*} \\clearpage There is a uniquely rich multi-wavelength dataset on the Coma cluster. X-ray observations covering a large area of the cluster have been made with ROSAT (White et al. 1993) and XMM-Newton (Briel et al.\\ 2001), showing the distribution and properties of the hot intra-cluster medium (ICM), and X-ray properties of individual galaxies have been studied by Finoguenov et al.\\ (2004) and by Hornschemeier et al. (2006). Coma has been shown by INTEGRAL to be an extended hard X-ray/soft $\\gamma$-ray source (Renaud et al.\\ 2006). At soft X-ray and Extreme Ultraviolet wavelengths there is a thermal excess (Lieu et al. 1996; Kaastra et al. 2003; Bonamente et al. 2003; Bowyer et al.\\ 2004). GALEX has been used to observe the cluster in the mid and near Ultraviolet and has sufficient spatial resolution to measure the UV properties of individual galaxies. In the infra-red, studies of the galaxies have been made using SPITZER, both with IRAC at 3.6 - 8 $\\mu$m (Jenkins et al.\\ 2007) and with MIPS at 24 and 70 $\\mu$m (Bai et al.\\ 2006). At radio wavelengths, VLA continuum maps cover much of the cluster (Miller et al.\\ in preparation), and in the HI 21 cm line there are extensive VLA imaging surveys (Bravo-Alfaro et al.\\ 2000, 2001), and single-dish spectra and fluxes for samples of spiral galaxies (Gavazzi et al.\\ 2006; Vogt et al.\\ 2004). This wealth of existing data makes Coma a prime target for studies of the origin and evolution of the galaxy content of clusters, and of its interaction with the other components (gas and dark matter). Moreover, as the richest and best studied local cluster, Coma is the zero-redshift baseline for many studies of high-redshift clusters (e.g. Jorgensen et al.\\ 2006). Comparison between low- and high-redshift clusters is vital for our understanding of their evolution, which in turn is essential if we are to disentangle evolutionary effects from the properties which tell us about their formation. We describe here the HST/ACS Coma Cluster Treasury survey, which aims to provide an unparalleled database of high spatial resolution images of a sample of cluster galaxies. At the distance of the Coma cluster ($\\sim$100 Mpc), the resolution of HST/ACS (0{\\arcsec}.1) corresponds to $\\sim$50 pc. This gives essentially the same physical resolution as ground-based observations have in Virgo and Fornax. Thus the HST/ACS Coma database provides a valuable comparison between high- and low-density clusters, for studies of the effect of environment on galaxy components. Whilst the HST observations are the prime data upon which this survey is based, it has already generated numerous ancillary observations with facilities such as Subaru, Keck, MMT, UKIRT. It is concurrent with surveys of the cluster in other wavebands, including the ultra-violet (GALEX), infra-red (SPITZER), X-ray (Chandra and XMM-Newton) and radio (VLA). ", "conclusions": "" }, "0801/0801.4316_arXiv.txt": { "abstract": "We report millimeter interferometric observations of polarized continuum and line emission from the massive star forming region G34.4. Polarized thermal dust emission at 3 mm wavelength and CO $J=1 \\rightarrow 0$ line emission were observed using the Berkeley-Illinois-Maryland Association (BIMA) array. Our results show a remarkably uniform polarization pattern in both dust and in CO J=$1 \\rightarrow 0$ emission. In addition, the line emission presents a consistent uniform polarization pattern over most of the velocity channel maps. These uniform polarization patterns are aligned with the north-south main axis of the filament between the main millimeter source (MM) and the ultra-compact H {\\scriptsize II} region, which are the central sources in G34.4, suggesting a magnetic field orthogonal to this axis. This morphology is consistent with a magnetically supported disk seen roughly edge-on. ", "introduction": "It is generally accepted that magnetic fields play an important role in the process of star formation; magnetic fields are involved in cloud support, fragmentation, and transfer of angular momentum. However, the magnetic field is the least observed physical quantity involved in such process. Magnetic field observations of molecular clouds are divided into measurements of the line-of-sight component of the magnetic field strength through the Zeeman effect and observations of the field in the plane of the sky through linear polarization of dust emission and spectral-line emission. The alignment of dust grains by a magnetic field is physically complicated and is still a matter of intense research. It is accepted, though, that aligned dust grains will produce polarized emission perpendicular to the projection of the magnetic field onto the plane of the sky. For a recent review of alignment theories see \\citet{Lazarian2007}. Spectral-line linear polarization has been suggested to arise from molecular clouds under anisotropic conditions, like large velocity gradients (or LVGs) \\citep{Goldreich1981}. The prediction suggests that a few percent of linearly polarized radiation should be detected from molecular clouds and circumstellar envelopes in the presence of a magnetic field. This polarization will be either parallel or perpendicular to the projection of the field onto the plane of the sky. To obtain a qualitative understanding about this effect, consider the CO molecule emitting unpolarized radiation. Under a magnetic field, a CO molecule will develop a small splitting in its rotational, $J$, energy levels. These magnetic sub-levels will produce radiation components labeled $\\sigma$ for the $\\left| M-M \\acute{} \\right| $ $=1$ transitions, and $\\pi$ for the $\\left| M-M \\acute{} \\right| $ $=0$ transitions, where both components can be linearly polarized either perpendicular or parallel to the magnetic field. If the gas behaves under isotropic conditions (e.g. no velocity gradients) for any direction the $\\sigma$ and $\\pi$ will populate equally, so the radiation components emerging from the radiative decays of these states will combine to give zero net polarization. Now on the other hand, large velocity gradients present in molecular clouds will produce anisotropies in the optical depths for the CO molecular transitions at different directions. If the velocity gradients are smaller in directions parallel to the magnetic field than in those perpendicular to the field, the optical depths parallel to the field will be larger than the optical depths perpendicular to them. Therefore, the escape of radiation involved in de-exciting the upper $J$ state will be then reduced more in directions along the field lines than in directions perpendicular to them, which will lead to populations of the magnetic $\\sigma$ sub-states that are larger than the populations of the $\\pi$ sub-states due to the difference in the angular distributions of both radiation components. The angular distribution of the $\\sigma$ component peaks in directions along the field lines whereas the $\\pi$ component peaks in directions perpendicular to the magnetic field. In this way, the rate of de-excitation for $\\sigma$ will have a larger decrease, due to photon trapping, than the rate of de-excitation for the $\\pi$ radiation component. Because, in this picture, the $\\sigma$ component will have a larger population, its emission will be stronger relative to the $\\pi$ component giving raise to a small amount of linear polarization in the CO emission with the polarization of the $\\sigma$ component, or perpendicular to the magnetic field. The effect was first detected by \\citet{Glenn1997} for the CS molecules while \\citet{Greaves1999} detected CO polarized emission for $(J=2 \\rightarrow 1)$ and $(J=3 \\rightarrow 2)$ transitions. In order to efficiently map polarized emission and infer detailed information about the magnetic field morphology, high resolution observations are required. The BIMA millimeter interferometer has been used previously to obtain high-resolution polarization maps in several star forming cores \\citep{Lai2001,Lai2002, Lai2003,Cortes2005,Cortes2006a,Cortes2006b}. These previous results show fairly uniform polarization morphologies over the main continuum sources, suggesting that magnetic fields are strong, and therefore cannot be ignored by star formation theory. However, the number of star formation regions with maps of magnetic fields remains small, and every new result is statistically significant. In this work we present polarization maps of the massive star forming region G34.4, obtained with the BIMA array. We measured continuum polarization at 3 mm and CO $J=1\\rightarrow 0$ line polarization, obtaining interferometric maps for both line and continuum. The remainder of this paper is divided into five sections. Section 2 reviews information about the source, section 3 describes the observation procedure, section 4 presents the results, section 5 gives the discussion, and section 6 the conclusions and summary. ", "conclusions": "The G34.4+0.23 MM massive star forming core was observed in the 3 mm band in polarized continuum and in CO $J=1 \\rightarrow 0$ polarized line emission with the BIMA array. The data show a uniform polarization pattern in both emissions. The P.A. obtained from the continuum data has an average value of $<\\phi> = -8^{\\circ} \\pm 5^{\\circ}$, and from the CO polarization $<\\phi> = -2^{\\circ} \\pm 8^{\\circ}$. These results suggest a magnetic field perpendicular to the main axis of the filament, between the MM source and the UC H {\\scriptsize II} region, in G34.4. The morphology suggests a flattened disk with the magnetic field along the minor axis, as predicted by the theory of magnetically supported molecular clouds. From our 3 mm continuum observations, we estimate a total core mass of 400 $M_{\\sun}$, in agreement with previous observations. Both line and continuum emission agrees in morphology with previous work by \\citet{Shepherd2004,Shepherd2007}. Finally, additional observations, particularly higher resolution interferometric mapping along the filament (including the most northern source seen in Figure \\ref{1}), will help to constrain, in greater detail, the morphology of the field and to obtain estimates of its strength in the plane of the sky. P. C. Cortes acknowledges support from the ALMA-CONICYT fund for development of Chilean Astronomy through grant 31050003. P. C. Cortes would also like to acknowledge the support given by NCSA and the Laboratory for Astronomical Imaging at University of Illinois at Urbana-Champaign during this research. Finally, P. C. Cortes would like to acknowledge the contribution by Patricio Sanhueza in making Figure \\ref{1}. R. M. Crutcher acknowledges support from NSF grants AST 05-40459 and 06-06822. L. Bronfman acknowledges support from the Chilean Center for Astrophysics FONDAP 15010003. \\altaffiltext{1}{National Radio Astronomy Observatory, P.O. Box O, 1003 Lopezville Rd, Socorro, NM 87801.} \\altaffiltext{2}{The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.}" }, "0801/0801.4299_arXiv.txt": { "abstract": "{The coming Planck and Herschel missions will survey the sky at unprecedented angular scales and sensitivities. Simulations are needed for better interpretating the results of the surveys and for testing new methods of, e.g., source extraction and component separation.} {We present new simulations of the infrared and submillimeter cosmic background, including the correlation between infrared galaxies. The simulations were used to quantify the source-detection thresholds for Herschel/SPIRE and Planck/HFI, as well as to study the detectability of the cosmic infrared background correlated fluctuations.} {The simulations are based on an empirical model of IR galaxy evolution. For these correlations, we only included the linear clustering, assuming that infrared galaxies are biased tracers of the dark-matter fluctuation density field.} {We used the simulations with different bias parameters to predict the confusion noise for Herschel/SPIRE and Planck/HFI and the completeness levels. We also discuss the detectability of the linear clustering in Planck/HFI power spectra, including the foreground and backgrounds components.} {Simulated maps and catalogs are publicly available online at http://www.ias.u-psud.fr/irgalaxies/simulations.php. } ", "introduction": "The cosmic infrared background (CIB) ($\\lambda\\geq8\\mu m$) is the relic emission of the formation and evolution of galaxies. The first observational evidence of this background was reported by \\citet{1996A&A...308L...5P} and then confirmed by \\citet{1998ApJ...508...25H} and \\citet{1998ApJ...508..123F}. The discovery of a surprisingly high amount of energy in the CIB has shown the importance of studying its sources to understand how the bulk of stars was formed in the Universe. Deep cosmological surveys have been carried out thanks to ISO \\citep [see] [for reviews] {2000ARA&A..38..761G,2005SSRv..119...93E} mainly at 15 $\\mu m$ with ISOCAM \\citep [e.g.] [] {2002A&A...384..848E}; at 90 and 170 $\\mu m$ with ISOPHOT \\citep [e.g.] [] {2001A&A...372..364D}; to SPITZER at 24, 70, and 160 $\\mu m$ \\citep [e.g.] [] {2004ApJS..154...70P,2004ApJS..154...87D} and to ground-based instruments such as SCUBA \\citep [e.g.] [] {1998astro.ph..9121H} and MAMBO \\citep [e.g.] []{2000astro.ph.10553B} at 850 and 1300 $\\mu m$, respectively. These surveys have allowed for a better understanding of the CIB and its sources \\citep [see] [for a general review] {2005ARA&A..43..727L}. Some of the results include: the energy of the CIB is dominated by starbursts although AGN (active galactic nucleus) contribute too, and the dominant contributors to the energy output are the LIRGs (luminous IR galaxies) at $z\\sim 1$ and ULIRGs (ultra luminous IR galaxies) at $z\\sim 2-3$. \\\\ Determination of the CIB by the COBE satellite has been hindered by the accuracy of subtracting the foreground by only providing just upper limits at 12, 25, and 60 $\\mu m$ \\citep {1998ApJ...508...25H}, lower limit has been derived at 24 $\\mu m$ by \\citet{2004ApJS..154...70P} as well as the contribution of 24 $\\mu m$ galaxies to the background at 70 and 160 $\\mu m$ \\citep {2006A&A...451..417D}. The contribution of the galaxies down to 60 $\\mu Jy$ at 24 $\\mu m$ is at least 79\\% of the 24 $\\mu m$ background, and 80\\% of the 70 and 160 $\\mu m$ background. For longer wavelengths, recent studies have investigated the contribution of populations selected in the near-IR to the far-infrared background (FIRB, $\\lambda > 200 \\mu m$): 3.6 $\\mu m$ selected sources to the 850 $\\mu m$ background \\citep {2006ApJ...647...74W} and 8 $\\mu m$ and 24 $\\mu m$ selected sources to the 850 $\\mu m$ and 450 $\\mu m$ backgrounds \\citep{2006ApJ...644..769D}. Similar studies with Planck and Herschel will provide even more evidence of the nature of the FIRB sources.\\\\ Studying correlations in the spatial distribution of IR galaxies as a function of redshift is an essential observation (parallel to the studies of individual high-redshift, infrared, luminous galaxies), to understand the underlying scenario and physics of galaxy formation and evolution. A first study has been done using the 850 $\\mu$m galaxies \\citep{2004ApJ...611..725B}. Although the number of sources is quite small, they find evidence that submillimiter galaxies are linked to the formation of massive galaxies in dense environments destined to become rich clusters. This has now been directly supported by the detection of the clustering of high-redshift 24 $\\mu$m selected ULIRGs and HyperLIRGs \\citep{2006ApJ...641L..17F,2007MNRAS.375.1121M}. Studying correlations with individual IR galaxies is very hard due to either high confusion noises, instrumental noises, or small fields of observation. It has been shown that the IR-background anisotropies could provide information on the correlation between the sources of the CIB and dark matter for large-scale structures \\citep [] [hereafter HK] {2001ApJ...550....7K,2000ApJ...530..124H} and on the large-scale structure evolution. First studies at long wavelengths have only detected the shot-noise component of the fluctuations: \\citet{2000A&A...355...17L} at 170 $\\mu m$, \\citet{2000A&A...361..407M} at 90 and 170 $\\mu m$, \\citet{2001phso.conf..471M} at 60 and 100 $\\mu m$. \\citet{2007ApJ...665L..89L} and \\citet{2007A&A..4512G} report first detections of the correlated component using Spitzer/MIPS data at 160 $\\mu m$. \\citet{2007ApJ...665L..89L} measured a linear bias $b \\sim 1.7$. \\\\ Future observations by Herschel and Planck will allow us to probe the clustering of IR and submm galaxies. Nevertheless these experiments will be limited, the confusion and instrumental noises will hinder detections of faint individual galaxies. Clustering thus has to be analysed in the background fluctuations \\citep [e.g.] [] {2007astro.ph..3210N}. The need for a prior understanding of what could be done by these experiments has motivated us to develop a set of realistic simulations of the IR and sub-mm sky. \\\\ In Sect. 2 we present the model on which are based our simulations. In Sect. 3 we discuss how the simulations are done and present a set of simulated sky maps and their corresponding catalogs. Different catalogs are created for 3 different levels of correlation between the IR galaxy emissivity and the dark-matter fluctuation density field (strong, medium, and no correlation). For each of these catalogs, we can create maps of the sky at any given IR wavelength and simulate how different instruments will see them. We focus in this paper on Planck/HFI and Herschel/SPIRE. In Sect. 4 we use the simulated maps to give predictions for the confusion noise, the completeness, and the detection limits for each of the study cases, including the instrumental noise. In Sect. 5 we present the power spectra of the CIB anisotropies for Planck/HFI and discuss their detectability against the significant sources of contamination (shot noise, cirrus, and cosmic microwave background (CMB)). \\\\ Throughout the paper the cosmological parameters were set to $h=0.71,\\Omega_{\\Lambda}=0.7,\\Omega_{m}=0.27$. For the dark-matter linear clustering we set the normalization to $\\sigma_{8}=0.8$. ", "conclusions": "This paper presented new simulations of the cosmic infrared and submillimeter background. The simulations are based on an empirical model of IR galaxy evolution (the LDP model) combined with a simple description of the correlation. The IR galaxy spatial distribution follows that of the dark-matter halo density field, with a bias parameter accounting for the possibility of the luminous matter being more or less correlated with the dark matter. The simulated maps and their catalogs are publicly available at http://www.ias.u-psud.fr/irgalaxies/simulations.php. Other maps are available upon request. These maps are intended to be a useful tool for planning future large IR and submillimeter surveys. In this paper, we used the maps to predict the confusion noise and completeness levels for Planck/HFI and Herschel/SPIRE. We also predicted the power spectra of correlated CFIRB anisotropies for Planck/HFI which will be a major advance in the study of the CFIRB anisotropies at large scales (i.e. $\\ell<2000-5000$ depending on the wavelengths). Further analysis of the CFIRB anisotropies, including the use of stacking analysis to isolate the anisotropies in different redshift ranges, will be presented in a second paper, now in preparation." }, "0801/0801.2467.txt": { "abstract": "We present results on integral-field optical spectroscopy of five luminous Blue Compact Dwarf galaxies. The data were obtained using the fiber system \\textsc{integral} attached at the William Herschel telescope. The galaxies Mrk~370, Mrk~35, Mrk~297, Mrk~314 and III~Zw~102 were observed. The central $33\\farcs6\\times29\\farcs4$ regions of the galaxies were mapped with a spatial resolution of $2\\farcs7$~spaxel$^{-1}$, except for Mrk~314, in which we observed the central $16\\arcsec\\times12\\arcsec$ region with a resolution of $0\\farcs9$~spaxel$^{-1}$. We use high-resolution optical images to isolate the star-forming knots in the objects; line ratios, electron densities and oxygen abundances in each of these regions are computed. We build continuum and emission-line intensity maps as well as maps of the most relevant line ratios: [\\ion{O}{3}]~$\\lambda5007$/\\Hb, [\\ion{N}{2}]~$\\lambda6584$/\\Ha, and \\Ha/\\Hb, which allow us to obtain spatial information on the ionization structure and mechanisms. We also derive the gas velocity field from the \\Ha\\ and [\\ion{O}{3}]~$\\lambda5007$ emission lines. We find that all the five galaxies are in the high end of the metallicity range of Blue Compact Dwarf galaxies, with oxygen abundances varying from $Z_{\\sun}\\simeq0.3$ to $Z_{\\sun}\\simeq1.5$. The objects show \\ion{H}{2}-like ionization in the whole field of view, except the outer regions of III~Zw~102 whose large [\\ion{N}{2}]~$\\lambda6584$/\\Ha\\ values suggest the presence of shocks. The five galaxies display inhomogeneous extinction patterns, and three of them have high \\Ha/\\Hb\\ ratios, indicative of a large dust content; all galaxies display complex, irregular velocity fields in their inner regions. ", "introduction": "Blue Compact Dwarf (BCD) galaxies are low-luminosity and compact objects, which show optical spectra similar to those presented by \\ion{H}{2} regions in spiral galaxies \\citep{ThuanMartin81}. They are also low-metallicity systems \\citep{Searle72,Lequeux79,KunthSargent83,Masegosa94} with intense star formation activity (rates ranging between 0.1 and $1\\;M_{\\sun}$ yr$^{-1}$; \\citealt{Fanelli88}). In most BCDs the gas consumption time scale is much shorter than the age of the Universe, which, together with their low metal content, raised initially the hypothesis that they could be truly young galaxies, forming their first generation of stars \\citep{Searle72,Kunth88}. Nowadays it is clear that the great majority of them are not young, but rather old systems (\\citealt{Papaderos96a}; \\citealt[][=C01a]{Cairos01a}; \\citealt[][=C01b]{Cairos01b}; \\citealt{BergvallOstlin02}; \\citealt[][=C03]{Cairos03}; \\citealt{Gildepaz03}) undergoing short starbursts followed by longer quiescent periods. An essential requisite to comprehend the formation and evolution of BCDs, and to elaborate coherent pictures of dwarf galaxy evolution, is to understand the process of star-formation in these galaxies ---~how their starburst activity ignites and how it propagates afterward~---, and to derive their star-forming (SF) histories. Even though both issues have been the subject of a considerable observational and interpretative effort during the last years, they are still far from being well understood. Different scenarios have been invoked to explain the onset of star formation in these galaxies. Some of them favored internal processes, such as the Stochastic Self-Propagating Star Formation proposed by \\cite{Gerola80}, or the hypothesis of cyclic gas re-processing, that is, the cyclic expulsion and later accretion of the Interstellar Medium \\citep{Davies88}. Alternatively, interactions and/or mergers have been proposed as the mechanism responsible for triggering the star formation in dwarfs. From optical surveys \\citep{Campos93,TellesTerlevich95,TellesMaddox00} it was first concluded that gravitational interactions with optically bright galaxies are too rare to play a relevant role. However, studies at radio wavelengths \\citep{Taylor93,Taylor95,Taylor96} established that \\ion{H}{2}/BCDs do have companions, most of them gas-rich, faint objects, which may go undetected in the optical. In agreement with these findings, more recent optical studies that extended the searches at lower masses and luminosities also found that a substantial fraction of SF dwarfs possess low mass companion galaxies \\citep{Noeske01,Pustilnik01}. \\cite{Ostlin01} and \\cite{BergvallOstlin02} put forward the idea that a merger between two galaxies with different metallicities, one gas-rich and one gas-poor, or infall of intergalactic clouds, could be the most plausible explanation for the starburst activity, at least in the most luminous BCDs. Recent studies, focused on individual objects, have shown that interactions do play a decisive role in the evolution of these systems \\citep{Johnson04,BravoAlfaro04,BravoAlfaro06}. Much work must still be done to elucidate the star formation histories of BCDs. The first step is to derive the properties of their stellar populations, quite a difficult task in such complex objects. The great majority of them cannot be resolved into stars, and the only way to study their stellar populations is by comparing their integrated properties with the predictions of evolutionary synthesis models. At any location in the galaxy, the emitted flux is the sum of the emission from the local starburst, the flux produced by the nebula surrounding the young stars, and the emission from the underlying, old stellar population, all possibly modulated by dust (\\citealt{Cairos00}; \\citealt[][=C02]{Cairos02}; \\citealt[][=C07]{Cairos07}). Another issue that was recently brought into discussion is the possible heterogeneity of the BCD class. Among the galaxies classified as BCDs we find a wide variety of objects spanning a large interval in luminosities ($-21.5 < M_{B} < -14.0$) and morphologies. \\cite{LooseThuan86} developed a classification scheme for BCDs based on the appearance of the starburst and the shape of the external envelopes, distinguishing four subclasses: \\textit{iE} galaxies, which show a complex inner structure with several SF regions over-imposed on an outer regular envelope; \\textit{nE}, objects with a clearly defined nucleus and regular isophotes; \\textit{iI}, which have irregular outer and inner parts and, finally, \\textit{iO} galaxies, in which an outer envelope is not found. C01b have grouped the BCDs in four categories according to the position and morphology of the SF knots: {\\em nuclear starburst}, which are galaxies with a prominent central starburst; {\\em extended starburst}, galaxies with star formation spread over the entire galaxy, {\\em chain starburst}, objects in which the SF knots are aligned to form a chain and finally, {\\em cometary starburst}, galaxies with a \"cometary\" appearance, that is the star formation concentrated in one side of the galaxy. This finding introduces an additional complication, because it opens the possibility that different star formation mechanisms may operate in BCDs and that galaxies classified as BCDs may actually have different star formation histories and evolutionary paths. Spectrophotometric studies that put together high quality data in a large wavelength range are fundamental to get insights into the star-formation process and history of BCDs. However, very few such studies have been carried out so far, and all of them focused on individual galaxies (\\citealt{Papaderos99,GildePaz00}; C02; \\citealt{Guseva03a,Guseva03b,Guseva03c}; C07). These studies rely on conventional observing techniques: they combine good resolution broad-band frames in different bands (to spatially isolate the different stellar populations and map the dust distribution), with narrow-band imaging (needed to map the gas distribution, isolate the starburst regions and to derive their physical properties) and spectroscopy data (to derive the internal extinction, compute physical parameters and chemical abundances of the gas, and remove the contribution from emission lines); a sequence of long-slit spectra sweeping the regions of interest is usually taken. Therefore, these analyses require a great amount of observing time; besides, varying instrumental and atmospheric conditions make combining all the data together quite complicated. Long-slit spectroscopy has the additional problem of the uncertainty on the exact location of the slit. We have thus undertaken a project to carry out an Integral Field Spectroscopy (IFS) mapping of a large sample of BCD galaxies. IFS is the ideal observational technique to study BCDs: each single exposure contains both spatial and spectral information in a large area of the galaxy, so just in one shot we collect information for all the SF regions as well as for the low surface brightness (LSB) stellar component. Besides, the kinematical information also allows us to investigate what mechanisms ignite the star-formation in BCDs. In terms of observing time, IFS observations of BCDs are an order of magnitude more efficient than traditional observing techniques, and provide simultaneous data for all spatial resolution elements under the same instrumental and atmospheric conditions, which guarantees a greater homogeneity of the dataset. Here we present the first results from this project: we study five galaxies representative of the group of BCDs populating the high luminosity and metallicity range, and having rather complex morphologies. These objects, often referred to as Luminous Blue Compact galaxies (LBCGs, see \\citealt{Ostlin98,Kunth00}; C01b) have a special significance in the general scenario of galaxy formation and evolution, as they are claimed to be the local counterparts of the luminous, compact, SF galaxies detected at higher redshifts \\citep{Garland04}. The proximity of these systems allows us to study their structure, stellar populations, star formation processes and chemical abundances with an accuracy and spatial resolution that cannot be achieved at intermediate/high redshift. ", "conclusions": "We present results from what is, as far as we know, the first Integral Field Spectroscopy analysis of a sample of galaxies catalogued as Blue Compact Dwarfs. % % Esto lo quite: \"From this study we highlight the following results:\" % \\begin{itemize} \\item With the help of high resolution optical images, we define several regions of interest in each object, and for each of them we measure emission line fluxes and compute the most relevant line ratios. These data show that: \\begin{itemize} \\item The strength of emission lines and absorption features, the shape of the continuum, and the presence of other relevant lines significantly vary across each galaxy. This indicates varying physical conditions and/or stellar content. \\item All the regions studied have \\ion{H}{2}-like ionization (i.e. star-formation). \\item All the identified regions in the observed galaxies have low electron densities, ranging from $\\leq 100$ to 360 cm$^{-3}$, typical of classical \\ion{H}{2} regions. The electron density also shows significant spatial variations in four out of the five galaxies studied. \\item The derived oxygen abundances are relatively large in all the cases, ranging from $Z_{\\sun}\\simeq0.3$ to $Z_{\\sun}\\simeq1.5$. No significant oxygen abundances variations within a same object are found. \\end{itemize} \\item We build maps of the stellar continuum, emission line fluxes and excitation ratios, finding that: \\begin{itemize} \\item Continuum and emission line morphologies are generally different. \\item The excitation ratios [\\ion{O}{3}]~$\\lambda5007$/\\Hb\\ and [\\ion{N}{2}]~$\\lambda6584$/\\Ha\\ are typical of HII regions in the whole observed fields. Only in the outer regions of III~Zw~102 the [\\ion{N}{2}]~$\\lambda6584$/\\Ha\\ ratio may suggest the presence of shocks. \\item All the galaxies present a complex extinction pattern. Assuming that the extinction coefficient is constant across the whole galaxy can lead to considerable errors in the derivation of magnitudes and colors and in the determination of the star formation rate and ages. \\item The \\Ha/\\Hb\\ ratios are higher that the theoretical values, indicating the presence of significant amounts of dust in these galaxies. \\end{itemize} \\item In all the five galaxies the central regions display complex, distorted ionized gas velocity fields, though large scale ordered motions are present in three of them. With the current data we cannot determine whether these perturbed velocity fields are the signature of interaction/mergers episodes or they are the result of shocks, stellar winds or supernovas. Further IFS observations are essential to investigate this issue: higher resolution spectra, in order to measure the velocity dispersion of the gas, as well as deeeper observations which allow to map the kinematics of the stars, are indeed fundamental. \\item The galaxies studied here, although classified as BCDs, all show characteristics very different from those of \"genuine\" BCD galaxies: they have relatively high metallicities and significant amounts of dust; they show also variations of such properties as electron densities, extinction and ionization degrees, across the galaxies. These properties make them an especially attractive area, as they could be the local counterparts of the blue, high metallicity, vigorously starbursting galaxies, detected at intermediate redshift. \\end{itemize}" }, "0801/0801.3646_arXiv.txt": { "abstract": "We perform numerical simulations of solid particle motion in a shearing box model of a protoplanetary disc. The accretion flow is turbulent due to the action of the magnetorotational instability. Aerodynamic drag on the particles is modelled using the Epstein law with the gas velocity interpolated to the particle position. The effect of the magnetohydrodynamic turbulence on particle velocity dispersions is quantified for solids of different stopping times $t_{\\rm{s}}$, or equivalently, different sizes. The anisotropy of the turbulence is reflected upon the dispersions of the particle velocity components, with the radial component larger than both the azimuthal and vertical components for particles larger than $\\sim $ 10 cm (assuming minimum-mass solar nebula conditions at 5 AU). The dispersion of the particle velocity magnitude, as well as that of the radial and azimuthal components, as functions of stopping time, agree with previous analytical results for isotropic turbulence. The relative speed between pairs of particles with the same value of $t_{\\rm{s}}$ decays faster with decreasing separation than in the case of solids with different stopping time. Correlations in the particle number density introduce a non-uniform spatial distribution of solids in the 10 to 100 cm size range. Any clump of particles is disrupted by the turbulence in less than one tenth of an orbital period, and the maximally concentrated clumps are stable against self-gravitational collapse. ", "introduction": "In order for planetesimals to form in a circumstellar disc, the parent dust grains must meet a variety of kinematic conditions if they are to grow in mass and size. The formation of kilometre-sized solid bodies by gravitational instability requires, through the Toomre criterion, that the dust spatial density be $\\rho_{d}>\\Omega^{2}/\\pi G$, where $\\Omega$ is the disc angular velocity and $G$ the gravitational constant. Such densities could be achieved in a mid-plane dust layer after sedimentation has occurred. On the other hand, growth by collisional sticking may be accomplished if the grain relative speeds acquire certain values: experiments have shown that, for micron-sized grains, collision speeds less than $\\sim$0.2 m/s lead to aggregates with a fractal geometry (Blum \\& Wurm 2000), and above $\\sim$ 1 m/s disruption of the colliding agglomerates takes place. While binary collisions of the smallest dust grains are caused by Brownian motion, other sources of relative velocities in a protoplanetary nebula may come into play as the interaction between solids and gas decreases. Differential sedimentation, radial drift and turbulence can provide the necessary kinetic energies to drive the collisional motion of solids, in addition to changing their spatial distribution throughout the nebular disc. In this paper we focus on the effect of a turbulent gas flow on solid particle velocities and spatial arrangement. The formalism of V\\\"{o}lk et al. (1980) makes it possible to calculate the velocity dispersions of particles in isotropic turbulence. The effect of the fluctuating gas velocity on individual solids is approximated by a sum over two types of turbulent eddies: those with spatial frequencies $k$ and turnover times $t_{k}>t_{\\rm{s}}$, where $t_{\\rm{s}}$ is the particle stopping time, and which merely advect the particle; and those with $t_{k}0\\,. \\end{equation} In this case, the coincidence problem is reduced to a simple choice of parameters to match $ \\Omega_{\\text{dark energy}}/ \\Omega_{\\text{dark matter}}$ to observations. Since the accelerated scaling solution is an attractor, no fine-tuning of initial conditions is needed. However, the dynamics that produces such scaling in the dark sector may have other undesirable consequences. Here we study quintessence with an exponential potential, \\begin{equation} V(\\varphi)= V_0 \\exp \\left(-\\kappa \\lambda \\varphi\\right)\\,,~~ \\kappa^2:=8\\pi G\\,, \\end{equation} where $\\lambda$ is dimensionless and $V_0>0$. The dynamics of a universe with exponential quintessence and an uncoupled perfect fluid have been studied, and these models do not admit late-time accelerated scaling attractors that satisfy Eq.~(\\ref{scal})~\\cite{quin}. When we introduce a coupling between the quintessence and dark matter, accelerated scaling attractors are possible~\\cite{Wetterich:1994bg,Amendola:1999qq}. However, in some models, this is achieved at the expense of introducing other problems which can rule out the model~\\cite{Amendola:2006qi}. A general coupling between a quintessence field $\\varphi$ and dark matter (with density $\\rho_c$) may be described in the background by the balance equations, \\begin{eqnarray} \\label{relation} \\dot \\rho_c &=& - 3H\\rho_c-Q\\,,\\label{cc}\\\\ \\dot \\rho_{\\varphi} &=& - 3H(1+w_{\\varphi})\\rho_{\\varphi}+ Q\\,. \\label{kg1} \\end{eqnarray} Here $Q$ is the rate of energy density exchange in the dark sector, and \\be \\label{sq} Q~\\left\\{ \\begin{array}{l} >0\\\\ <0 \\end{array} \\right. ~~ \\Rightarrow ~~ \\mbox{energy transfer}~\\left\\{ \\begin{array}{l} \\mbox{dark matter $\\to$ dark energy}\\\\ \\mbox{dark energy $\\to$ dark matter}\\end{array} \\right. \\ee The dark energy equation of state parameter is \\be w_\\varphi:={ p_\\varphi \\over \\rho_\\varphi}={{1\\over 2}\\dot\\varphi^2-V(\\varphi) \\over {1\\over 2}\\dot\\varphi^2+V(\\varphi)}\\,.\\label{wp} \\ee The modified Klein-Gordon equation follows from Eq.~(\\ref{kg1}): \\begin{equation} \\ddot \\varphi +3H \\dot \\varphi + \\frac{dV}{d\\varphi}= {Q \\over \\dot\\varphi}\\,. \\label{kg} \\end{equation} When we include baryons ($ \\rho_b$) and radiation ($\\rho_r $), the remaining evolution equations are \\begin{eqnarray} \\dot \\rho_b &=& - 3H\\rho_b \\,,\\\\ \\dot {\\rho}_r &=& - 4H\\rho_r \\,,\\\\ \\dot H &=& -\\frac{\\kappa^2}{2}\\left[\\rho_c+\\rho_b+ \\frac{4}{3}\\rho_r+\\dot\\varphi^2 \\right], \\label{ray} \\end{eqnarray} subject to the Friedman constraint, \\begin{eqnarray} \\label{fried} \\Omega_c+\\Omega_b+\\Omega_r +\\Omega_{\\varphi}=1\\,,~~ \\Omega := \\frac{\\kappa^2 \\rho}{3H^2}. \\end{eqnarray} We can define effective equation of state parameters for the dark sector, which describe the equivalent uncoupled model in the background: $\\dot \\rho_c +3H(1+ \\wc)\\rho_c=0 $, $\\dot \\rho_\\varphi +3H(1+ \\wv)\\rho_\\varphi=0 $. By Eqs.~(\\ref{cc}) and (\\ref{kg1}), \\be \\label{weff} \\wc={Q \\over 3H\\rho_c}\\,,~~ \\wv= w_\\varphi -{Q \\over 3H \\rho_\\varphi}\\,. \\ee It follows that \\bea && Q >0 ~~ \\Rightarrow ~~ \\left\\{ \\begin{array}{ll} \\wc>0 & ~~\\mbox{dark matter redshifts faster than}~a^{-3}\\\\ \\wvw_\\varphi & ~~\\mbox{dark energy has less accelerating power} \\end{array} \\right. \\eea When $Q>0$ it is possible that $\\wv<-1$ (see~\\cite{Huey:2004qv} for specific examples). This means that the coupled quintessence behaves like a ``phantom\" uncoupled model -- but without any negative kinetic energies. Equations~(\\ref{cc})--(\\ref{ray}) are an autonomous system of the form \\begin{eqnarray} \\dot {\\mathbf{x}} = \\mathbf{f}(\\mathbf{x})\\,, \\label{math1} \\end{eqnarray} and the critical points satisfy $\\mathbf{f}(\\mathbf{x}_*)=0$. In order to study the stability of the critical points, we expand about them, $\\mathbf{x} = \\mathbf{x}_* +\\mathbf{u}$, and Eq.~(\\ref{math1}) yields \\begin{equation} \\dot {\\mathbf{u}} = {\\mathbf{f}'}(\\mathbf{x}_*) \\mathbf{u} + \\mathbf{g}(\\mathbf{x})\\,. \\end{equation} Here $\\mathbf{g}(\\mathbf{x})/||\\mathbf{x}|| \\rightarrow 0$ as $\\mathbf{x} \\rightarrow \\mathbf{x}_*$, and \\begin{equation}\\label{eigv} f'_{ij}(\\mathbf{x}_*) = \\frac{\\partial f_{i}}{\\partial x_{j}}(\\mathbf{x}_*)\\,, \\end{equation} is a constant non-singular matrix, whose eigenvalues encode the behaviour of the dynamical system near the critical point. If a component of $\\mathbf{f}$ can be written as a fraction $u(\\mathbf{x})/v(\\mathbf{x})$, then a critical point requires the vanishing of the numerator, $u(\\mathbf{x}_*)=0$. If the denominator also vanishes at the critical point, $v(\\mathbf{x}_*)=0$, then care is needed in obtaining the eigenvalues of the linearized system~(\\ref{math1}). Strictly speaking, the fraction $u(\\mathbf{x}_*)/v(\\mathbf{x}_*)$ may not be well defined. However, it is still possible to obtain analytical results via analysis of the behavior of the eigenvalues of $\\mathbf{f}'$ in the limit $v(\\mathbf{x}) \\rightarrow 0$. ", "conclusions": "\\label{sec:concl} We considered the background dynamics of a universe dominated by dark energy (in the form of exponential quintessence) and cold dark matter, where there is energy exchange in the dark sector, as in Eqs.~(\\ref{cc}) and (\\ref{kg1}), \\[ \\dot \\rho_c + 3H\\rho_c=-Q=-\\left[ \\dot \\rho_{\\varphi} + 3H(1+w_{\\varphi})\\rho_{\\varphi}\\right]. \\] For the previously introduced forms of $Q$, given in Eqs.~(\\ref{I}) and (\\ref{II}), \\[ \\mbox{(I): }~~ Q=\\sqrt{ 2 /3}\\, \\kappa\\,\\beta \\rho_c\\dot\\varphi\\,,~~~ \\mbox{(II): }~~ Q= \\alpha H \\rho_c\\,, \\] the phase space remains two-dimensional, as in the uncoupled case $Q=0$. We found the properties of the critical points for all signs of $\\lambda, \\alpha, \\beta$. The results are summarized in Tables~\\ref{crit} and \\ref{crit2}, and slightly extend previous work~\\cite{Amendola:1999qq,Holden:1999hm,Billyard:2000bh} in the case of pressure-free matter ($w_c=0$). In both models, critical point D is the cosmologically relevant point, because for nonzero coupling it includes accelerated scaling attractor solutions, \\[ 0<\\Omega_{c *}\\,,\\Omega_{\\varphi *}<1 ~\\mbox{ and }~ w_{\\text{tot} *} <-{1\\over3}\\,. \\] The stability behaviour of critical point D was investigated numerically, and is shown in Fig.~\\ref{DI} for model (I) and Fig.~\\ref{DII} for model (II). Our main results are for a new coupling model~\\cite{val}, defined in Eq.~(\\ref{III}), \\[ \\mbox{(III): }~~Q=\\Gamma \\rho_c\\,. \\] This has a similar form to model (II), but is more physical since the transfer rate $\\Gamma$ is determined only by local properties of the dark sector interaction at each event, and is not dependent on the universal expansion rate. When $\\Gamma>0$, this new model has the same form as simple models for the decay of dark matter particles to radiation~\\cite{Cen:2000xv}, or to quintessence~\\cite{Ziaeepour:2003qs}, and for the decay of the curvaton field into radiation~\\cite{Malik:2002jb}. Model (III) requires a three-dimensional phase space, since the Hubble rate cannot be eliminated from the equations for $x',y'$. This makes the dynamics more complicated than for models (I) and (II). In particular, a new set of late-time critical points arises in (III), and considerable analytical effort is required to identify these points and determine their stability properties. Our results are summarized in Tables~\\ref{crit3} and \\ref{eigen1}. We performed numerical integrations of the dynamical system in order to confirm the analytical results, and examples of these integrations are shown in Figs.~\\ref{lambda_one} and \\ref{lambda_four}. The cosmologically relevant critical point is G, which allows for an accelerated critical solution (when $\\lambda^2<2$). However, this is not a scaling solution, since \\[ \\Omega_{c *}=0\\,,~~\\Omega_{\\varphi * }=1\\,, \\] which is similar to the asymptotic behaviour of the standard $\\Lambda$CDM model. This accelerated critical solution is an attractor when $\\gamma>0$, i.e., for the case when dark matter is decaying to dark energy. Note that in model (II), the decaying dark matter case, $\\alpha>0$, does not lead to any accelerated attractor (see Table~II). Model (III) with $\\Gamma>0$ produces an interesting class of models where dark matter decays to dark energy -- so that the primordial universe may have no dark energy -- and where this decay eventually leads to dark energy dominance, independent of initial conditions (since there is an attractor). Although such models do not solve the coincidence problem in the standard way, they may provide a new approach to the broader problem of explaining why dark energy dominates over dark matter only late in the universe's evolution. The background dynamics for coupling model (III) show new features not present in the previously investigated models (I) and (II). In order to confront this model with observations, the cosmological perturbations with a dark sector coupling of form (III) need to be investigated (see Ref.~\\cite{val}). \\[ \\] {\\bf Acknowledgements:}\\\\ We thank Luis Ure\\~na-L\\'opez, Elisabetta Majerotto, Luca Parisi, Israel Quir\\'os, Jussi V\\\"aliviita and Shinji Tsujikawa for useful discussions. GCC is supported by the Programme Alban, the European Union Programme of High Level Scholarships for Latin America, scholarship no.~E06D103604MX and the Mexican National Council for Science and Technology, CONACYT, scholarship no.~192680. RL is supported by the University of the Basque Country through research grant GIU06/37, and by the Spanish Ministry of Education and Culture through research grants FIS2004-01626 and FIS2004-0374-E. The work of RM is supported by STFC." }, "0801/0801.1279_arXiv.txt": { "abstract": "We compare the X-ray spectra and luminosities, in the 2-8 keV band, of known and suspected cataclysmic variables (CVs) in different environments, assessing the nature of these source populations. These objects include nearby CVs observed with ASCA; the Galactic Center X-ray source population identified by Muno et al.; and likely CVs identified in globular clusters. Both of the latter have been suggested to be dominated by magnetic CVs. We find that the brighter objects in both categories are likely to be magnetic CVs, but that the fainter objects are likely to include a substantial contribution from normal CVs. The strangely hard spectra observed from the Galactic Center sources reflect the high and variable extinction, which is significantly greater than the canonical $6\\times10^{22}$ cm$^{-2}$ over much of the region, and the magnetic nature of many of the brightest CVs. The total numbers of faint Galactic Center sources are compatible with expectations of the numbers of CVs in this field. ", "introduction": "The unprecedented spatial resolution of the {\\it Chandra X-ray Observatory} allows us to study populations of faint X-ray sources at distances of kiloparsecs. Large numbers of X-ray sources of moderate luminosities ($10^{31} 9.4$ km and $R > 7.8$ km (68\\% confidence) for PSRs J0030+0451 and J2124--3358, respectively. We explore the prospects of using future observatories such as \\textit{Constellation-X} and \\textit{XEUS} to conduct X-ray timing searches for MSPs not detectable at radio wavelengths due to unfavorable viewing geometry. We are also able to place strong constraints on the magnetic field evolution model proposed by Ruderman. The pulse profiles indicate that the magnetic field of an MSP does not have a tendency to align itself with the spin axis nor migrate towards one of the spin poles during the low-mass X-ray binary phase. ", "introduction": "Recent X-ray studies have revealed that a number of known rotation-powered millisecond pulsars (MSPs) exhibit predominantly thermal soft X-ray emission \\citep{Grind02,Zavlin06,Bog06a,Zavlin07}. The infered effective emission radii $R_{\\rm eff}$ indicate that this radiation is localized in regions on the neutron star (NS) surface much smaller than the expected stellar radius ($R_{\\rm eff}\\ll R$) but comparable to the classical radius of the pulsar magnetic polar cap $R_{pc}=(2\\pi R/cP)^{1/2}R$. This finding is in agreement with theoretical models of pulsars, which indicate that the conditions in the magnetosphere of a typical MSP favor heating of the polar caps to $\\sim$$10^6$ K by a return flow of energetic particles along the open magnetic field lines \\citep[see, e.g.,][for details]{Hard02,Zhang03}. As this heat is restricted to a small fraction of the NS, study of the X-ray spectra and pulse profiles of MSPs can reveal important information about the star such as the radiative properties of the NS surface, magnetic field geometry, and NS compactness ($R/R_S$, where $R_S=2GM/c^2$). This approach, originally proposed by \\citet{Pavlov97} in the context of radio MSPs, can serve as a valuable probe of key NS properties that are inaccessible by other observational means (e.g.~radio pulse timing). As shown by \\citet{Pavlov97}, \\citet{Zavlin98}, and \\citet{Bog07}, a model of polar cap thermal emission from an optically-thick hydrogen (H) atmosphere provides an excellent description of the X-ray pulse profiles of PSR J0437--4715, the nearest known MSP. On the other hand, a blackbody model is inconsistent with the X-ray timing data and can be definitively ruled out. Furthermore, there is compelling evidence for a magnetic dipole axis offset from the NS center \\citep{Bog07}. Finally, the compactness of PSR J0437--4715 is constrained to be $R/R_S>1.6$ (99.9\\% confidence). Thus, modeling of X-ray data of MSPs appears to be a very promising approach towards answering long-standing questions regarding the fundamental properties of MSPs and NSs, in general. The present paper represents an extension of the work described in \\citet{Bog07}. Herein we explore the detailed properties of the MSP X-ray emission model with particular emphasis on its use for constraints on the NS equation of state (EOS). The work is organized as follows. In \\S2 we examine the properties of our model; in \\S3 we discuss an application of our model to archival X-ray observations of PSRs J0030+0451 and J2124--3358. In \\S4 we present a discussion and end with conclusions in \\S5. \\begin{figure*}[t!] \\includegraphics[width=0.98\\textwidth]{f1.ps} \\caption{(\\textit{Left}) Representative synthetic hydrogen atmosphere lightcurves for a rotating $M=1.4$ M$_\\odot$, $R=10$ km NS with two antipodal hot spots for the four lightcurve classes (I-IV, from top to bottom, respectively), defined by Beloborodov (2002). The dashed lines show the individual flux contribution from the two hot spots while the solid line shows the total observed flux. All fluxes are normalized to the value corresponding to a face-on hot spot ($\\alpha=\\zeta=0$). Two rotational cycles are shown for clarity. (\\textit{Right}) Orthographic map projection of the MSP surface for the four pulse profiles. The dashed line shows the magnetic axis while the dotted line shows the line of sight to the observer. The hatched region corresponds to the portion of the star not visible to the observer.} \\end{figure*} ", "conclusions": "We have examined the properties of our model of thermal emission from hot spots on the surface of a neutron star covered by a hydrogen atmosphere, relevant for MSPs. Our investigation has demonstrated that energy-resolved modeling of the thermal X-ray pulse profiles and phase-resolved spectroscopy of the continuum emission can, in principle, be used to determine $M/R$ and the pulsar geometry to high accuracy, given observational data of sufficient quality and favorable values of $\\alpha$ and $\\zeta$. As shown in \\S2.1 the thermal pulse profiles are surprisingly insensitive to the details of the polar cap size and shape, the distance to the pulsar, and the uncertainty in the instrument effective area. This method represents the only feasible approach of studying MSP magnetic field, surface properties, and compactness, especially for isolated MSPs. Barring any deleterious effect such as additional hidden spectral components \\citep[see, e.g,][and references therein]{Bog06b} this approach can, in principle, lead to unprecedented insight into NS properties. Our model is found to be in agreement with the observed emission from the nearby solitary MSPs J0030+0451, and J2124--3358. As with PSR J0437--4715 \\citep{Bog07}, the relatively large pulsed fractions observed in PSRs J0030+0451 and J2124--3358 require the existence of a light-element atmosphere on the stellar surface and cannot be reproduced by a blackbody model for realistic NS radii. For J0030+0451 and J2124--3358 we are able to place interesting limits on the allowed compactness of $R>9.4$ km and $R>7.8$ km (68\\% confidence) assuming $M=1.4$ M$_{\\odot}$. Based on our findings in \\S2, we expect deeper observations of this MSP to lead to much tighter constraints on $M/R$, which in turn may firmly rule out certain families of NS EOS. The available thermal X-ray pulse profiles also provide useful constraints on magnetic field models of MSPs. Specifically, the positions of the magnetic polar caps on the NS surface implied by the X-ray data indicate that the magnetic field closely resembles the conventional oblique dipole model of pulsars rather than more exotic field configurations. The success of this approach towards elucidating crucial NS properties motivates further studies of the nearby sample of MSPs with both \\textit{Chandra} and \\textit{XMM-Newton}. Moreover, this makes MSPs particularly important targets for upcoming X-ray mission such as \\textit{Constellation-X} and \\textit{XEUS}. The great increase in throughput of these facilities will allow searches for new MSPs, detailed observations of a larger sample of known radio MSPs, and unprecedented constraints on key NS properties, especially the NS EOS." }, "0801/0801.4801_arXiv.txt": { "abstract": "We describe the construction of GROND, a 7-channel imager, primarily designed for rapid observations of gamma-ray burst afterglows. It allows simultaneous imaging in the Sloan $g'r'i'z'$ and near-infrared $JHK$ bands. GROND was commissioned at the MPI/ESO 2.2\\,m telescope at La Silla (Chile) in April 2007, and first results of its performance and calibration are presented. ", "introduction": "Simultaneous imaging in different filter-bands is of interest in a variety of astrophysical areas. The primary aim is to measure the spectral energy distribution or its evolution in variable objects in order to uncover the underlying emission mechanism. Examples are, among others, (1) monitoring of all kinds of variable stars (flare stars, cataclysmic variables, X-ray binaries) to determine the outburst mechanisms and differentiate between physical state changes and changes induced by geometrical variations, like eclipses; (2) monitoring of AGN to understand the physical origin of the observed variability; (3) determining the inclination of X-ray heated binaries \\cite{orosz} (4) mapping of galaxies to study the stellar population; (5) multi-color light curves of supernovae \\cite{tpm06}; (6) differentiate achromatic microlensing events \\cite{pac86} from other variables with similar light curves; (7) identifying objects with peculiar spectral energy distributions, e.g. photometric redshift surveys for high-$z$ active galactic nuclei, or identifying brown dwarfs; (8) follow-up observations of transiting extrasolar planets \\cite{jha00}; or (9) mapping of reflectance of solar system bodies as a function of their rotation to map their surface chemical composition \\cite{jew02}. A new need for multi-band imaging arose with the observation of a large number of gamma-ray burst (GRB) afterglows with the {\\it Swift} satellite \\cite{geh04}. With its much more sensitive instruments it detects GRBs over a very wide redshift range. Since intermediate to high-resolution spectroscopy to measure the physical conditions of the burst environment \\citep[e.g.][]{vre07} is constrained to the first few hours after a GRB explosion, a rapid determination of the redshift became important. This is best done with multi-band photometry (until integral field units have grown to several arcmin field-of-views) and deriving a photometric redshift based on the Ly$\\alpha$ break \\cite{lr00}. Previous and current instruments with simultaneous imaging capability in different filter bands include ANDICAM (A Novel Double-Imaging CAMera; two channels, one for visual, the other for near-infrared \\cite{dep98}, presently operated at a 1.3\\,m telescope), BUSCA (Bonn University Simultaneous CAmera; four visual channels \\cite{rei99}, presently operated at the 2.2\\,m telescope at Calar Alto), HIPO (High-speed Imaging optical Photometer for Occultations; two visual channels \\cite{dun04}, to be operated on SOFIA, the Stratospheric Observatory For Infrared Astronomy), MITSuME (Multicolor Imaging Telescopes for Survey and Monstrous Explosions; three channels with fixed bands g\\amin, R$_{\\rm C}$ and I$_{\\rm C}$ \\cite{kky07}, operated at a 50 cm telescope), TRISPEC (Triple Range Imager and Spectrograph; three channels with one CCD and two near-IR detectors and wheels for filters, grisms and Wollaston prisms \\citep{wny05}), SQIID (Simultaneous Quad Infrared Imaging Device; $JHK$ and narrow-band $L$ filters in front of individual 512X512 quadrants of an ALADDIN InSb array, designed for the f/15 Cassegrain foci of the KPNO 2.1-m and 4-m telescopes \\citep{edf92}), ULTRACAM (ULTRAfast, triple-beam CCD CAMera for high-speed astrophysics \\cite{dhi07}; portable instrument which has been used, among others, at the Very Large Telescopes at ESO, or the William Herschel Telescope, Canary Islands). Here we describe the design ($\\S$2) and performance ($\\S$6) of a 7-channel imager, called GROND ({\\bf G}amma-{\\bf R}ay Burst {\\bf O}ptical and {\\bf N}ear-Infrared {\\bf D}etector), which was specifically designed for GRB afterglow observations. We also mention some basics of the operation scheme ($\\S$3), related software ($\\S$4), and the changes to the telescope infrastructure which were implemented to use GROND for rapid follow-up observations ($\\S$5). ", "conclusions": "GROND is an imaging system capable of operating in seven colors simultaneously. It has been designed and built at MPE Garching, and commissioned at the 2.2\\,m MPI/ESO telescope on La Silla, Chile. First observations show that all properties are according to specifications/expectations. The first observations of gamma-ray bursts with GROND have also been obtained (Greiner \\etal\\ 2007a,b; Primak \\etal\\ 2007, Kr\\\"uhler \\etal\\ 2007). Fine tuning of the operations strategy as well as scheduling and analysis software in the upcoming weeks is expected to bring GROND into a fully operational condition, thus allowing the commencement of normal science operations. \\bigskip" }, "0801/0801.2616_arXiv.txt": { "abstract": "We present a preliminary analysis of ASTRO-F data of a complete sample of $\\sim$ 150 EROs (R-K$>$5) down to K$_{Vega}<$19, for which reliable photometric redshifts are available, in the range 0.8$<$z$<$2, selected over two fields (S7 and S2) of the MUNICS survey. The area covered is about 420 arcmin$^2$. We have imaged this area with AKARI telescope in N3 (3.4 $\\mu$m), N60 (65 $\\mu$m) and WL (150 $\\mu$m) down to 12 $\\mu$Jy in the N3 filter, in order to detect the rest frame H or K-band emission, thus providing an excellent sampling of the SED of our EROs. From a first analysis we have an identification rate of $\\sim$ 63\\% in the N3 filter over the S7 field. These data allow us to distinguish starburst from passive early type phenomena, to meseaure the SFR of the starburst component and to constrain the mass assembly of early type galaxies. ", "introduction": "While the general build-up of cosmic structures seems to be well described by hierarchical models of galaxy formation ($\\Lambda$CDM), the assembly of the baryonic mass on galactic scale still represents a weak point of the galaxy formation models. One of these difficulties consists in explaining the large population of Extremely Red Objects (R-K$>$5) EROs. This population is a mixture of early type galaxies (ETG) and dusty star forming galaxies (SFG). Hierarchical models should reproduce the abundance of massive red galaxies at 1$$100), bright (K$<$19) and complete sample of EROs. \\begin{figure} \\centering \\includegraphics[width = 6 cm, height = 5cm]{mignanoa_fig1.jpg} \\includegraphics[width = 7 cm, height = 5.3cm]{mignanoa_fig2.jpg} \\label{detection} \\caption{({\\it left panel}) The S7F5 field imaged with N3 filter (IRC camera). The yellow circle represent the EROs selected from the MUNICS catalogue. ({\\it right panel}) N3 counterpart distribution vs. r. Red histogram shows the counterparts included in the identification sample.} \\end{figure} ", "conclusions": "" }, "0801/0801.0878_arXiv.txt": { "abstract": "We present medium resolution spectropolarimetry and long term photo-polarimetry of two massive post-red supergiants, IRC~+10420 and HD~179821. The data provide new information on their circumstellar material as well as their evolution. In IRC~+10420, the polarization of the H$\\alpha$ line is different to that of the continuum, which indicates that the electron-scattering region is not spherically symmetric. The observed long term changes in the polarimetry can be associated with an axi-symmetric structure, along the short axis of the extended reflection nebulosity. Long term photometry reveals that the star increased in temperature until the mid-nineties, after which the photospheric flux in the optical levelled off. As the photometric changes are mostly probed in the red, they do not trace high stellar temperatures sensitively. And so, it is not obvious whether the star has halted its increase in temperature or not. For HD~179821 we find no polarization effects across any absorption or emission lines, but observe very large polarization changes of order 5\\% over 15 years. During the same period, the optical photometry displayed modest variability at the 0.2 magnitude level. This is unexpected, because large polarization changes are generally accompanied by strong photometric changes. Several explanations for this puzzling fact are discussed. Most of which, involving asymmetries in the circumstellar material, seem to fail as there is no evidence for the presence of hot, dusty material close to the star. A caveat is that the sparsely available near-infrared photometry could have missed periods of strong polarization activity. Alternatively, the variations can be explained by the presence of a non-radially pulsating photosphere. Changes in the photometry hint at an increase in temperature corresponding to a change through two spectral subclasses over the past ten years. ", "introduction": "Due to the steepness of the Initial Mass Function, massive stars ($\\ga$ 8 M$_{\\odot}$) are extremely rare. This is exacerbated by their comparatively short lifetimes. Yet, although rare, these objects have a crucial impact on the interstellar medium due to their strong winds and high mass-loss rates, and can dominate the light output of entire galaxies. An area of current interest is that massive evolved stars are often surrounded by bi-polar nebulae \\citep{weis:2003}. Later in the evolution of a star, using spectropolarimetry, it is now established that the ejecta of supernovae deviate from spherical symmetry (e.g. \\citealt{Wang_etal:2003,Leonard_etal:2005}). It is as yet unclear whether this is due to asymmetric explosions, axi-symmetric stellar winds or a pre-existing density contrast in the surrounding material (e.g. \\citealt{Dwarkadas_Balick:1998,Dwarkadas_Owocki:2002}). Wind-axisymmetry may imply fast rotation, and be related to the beaming of SN explosions, which may be the origin of the extremely luminous, beamed gamma-ray bursts (e.g. \\citealt{Meszaros:2003,Mazzali_etal:2003}). Here, we address the issue by investigating the circumstellar ejecta of two yellow hypergiants, IRC~+10420 and HD~179821. These objects are thought to have evolved off the post-Red Supergiant branch and are still surrounded by mass ejected during a previous mass losing phase. Only a few yellow hypergiants are known (see the review by \\citealt{de_Jager:1998}), and the number of such hypergiants with circumstellar dust is even smaller - only IRC~+10420 and HD~179821 belong to this class (see for example \\citealt{Oudmaijer08}). Therefore, the study of these two unique objects is important in its own right. As these stars are distant (3-5 kpc), the direct imaging of their innermost regions is currently beyond the reaches of current technology, although interferometry is starting to resolve the winds of evolved stars \\citep{dewit_etal:2008}. Observing with spectropolarimetry allows us to probe regions much closer to the star still. Spectropolarimetry was first effectively used in the study of classical Be stars using the presence of `line effects' \\citep{Poeckert:1975, Poeckert_Marlborough:1976}. These are changes in polarization across spectral lines that have an emission component. They occur because emission-line photons arise over a larger volume than the stellar continuum photons. Consequently, the emission-line photons undergo fewer scatterings as they `see' fewer electrons, resulting in a lower polarization than the continuum. We normally only observe a net polarization change if the geometry of the electron scattering region is aspherical. Many authors have confirmed that this technique provides evidence that envelope geometries around Be stars are indeed disk-like \\citep{dougherty_1992, Quirrenbach_etal:1997, Wood_Bjorkman_Bjorkman:1997}. More recently the technique has been used to investigate the geometry of circumstellar material around Herbig Ae/Be stars. Studies by \\cite{Oudmaijer_Drew:1999} and \\cite{Vink_etal:2002} show that most Herbig stars exhibit line effects, indicating aspherical electron-scattering regions. For a recent review, see \\cite{Oudmaijer:2007}. With regard to evolved stars, Davies et al. (2005) conducted a study of Luminous Blue Variables (LBVs) using spectropolarimetry. They found that 50$\\%$ of the objects observed exhibited polarization changes across H$\\alpha$, indicating that some asphericity lies at the base of the stellar wind. Furthermore, they found several objects for which the position angle varied randomly with time, leading them to conclude that the wind around these stars is clumpy (see also \\citealt{Nordsieck_etal:2001}). IRC~+10420 is now well accepted as a massive, evolved object (e.g. \\citealt{Jones_etal:1993, Oudmaijer_etal:1996} and \\citealt{Humphreys_etal:2002}). This is mainly based on its large distance, high outflow velocity (40 kms$^{-1}$) and high luminosity implied from the hypergiant spectrum. The situation for HD~179821 is less certain. The presence of non-radial pulsations coupled with comparatively modest photometric changes suggest a massive nature \\citep{LeCoroller_etal:2003}. Furthermore, its circumstellar material has a large expansion velocity of 30 km s$^{-1}$, as measured in CO, suggesting the star is a supergiant \\citep{Kastner_Weintraub:1995}. On the other hand, the overabundance of s-process elements and the low metallicity suggest HD~179821 is perhaps a lower mass post-AGB star \\citep{Zacs_etal:1996,Reddy_Hrivnak:1999,Thevenin_etal:2000}. Although the latter's nature is a bit more uncertain, there are some striking similarities between IRC~+10420 and HD~179821. When observed as part of a larger sample of post-AGB stars, these two objects are often markedly different from the rest. In particular, their high outflow velocities (the average outflow velocity for post-AGB and AGB stars is 15 kms$^{-1}$) require much higher luminosities if powered by radiation pressure alone \\citep{Habing_etal:1994}. They were the only objects that showed extensive reflection nebulae in a large survey by \\cite{Kastner_Weintraub:1995}. \\cite{Jura_etal:2001} point out the enormous difference between the space velocities of IRC~+10420 and HD 179821 when compared against low mass post-AGB stars. Furthermore, both objects have an exceptionally strong O{\\sc i} 7774 triplet absorption feature indicating a high luminosity (\\citealt{Humphreys_etal:1973} and \\citealt{Reddy_Hrivnak:1999}, based on \\citealt{slowik_1995}). We therefore proceed with both objects and assume they are evolved post-Red Supergiants. Recently, both objects have been observed at arcsecond resolution in CO by \\cite{Castro-Carrizo_etal:2007} who found, in accordance with previous estimates, that their mass loss rates exceeded 10$^{-4} \\, \\rm M_{\\odot} yr^{-1}$ when they were in the RSG phase. The envelopes show mild deviations from spherical symmetry in their data. This paper is organised as follows. In Section 2, we review the experimental setup and explain how the data has been reduced. We present our results for each object in Section 3, and use the new spectropolarimetric data together with past polarization measurements to investigate the nature of the circumstellar environments around each of the stars, which is discussed in Section 4. We conclude in Section 5. \\begin {table*} \\centering \\begin {tabular}{llcccccccccc} \\hline \\centering Object & Telescope & Date & Julian Date & $\\%$P & P.A. (Deg) & $\\%$P (H$\\alpha$) & P.A. (H$\\alpha$) (Deg) \\\\ \\hline HD~179821 \\\\ & AAT & 15/09/02 & 2452533 & 1.99 $\\pm$ 0.11 & 40 $\\pm$ 2 \\\\ & WHT & 30/09/04 & 2453279 & 2.00 $\\pm$ 0.15 & 36 $\\pm$ 3 \\\\ IRC~+10420 \\\\ & NOT & 16/01/98 & 2450830 & 1.95 $\\pm$ 0.05 & 174 $\\pm$ 1 \\\\ & NOT & 16/05/98 & 2450950 & 1.80 $\\pm$ 0.03 & 174 $\\pm$ 1 \\\\ & AAT & 18/09/02 & 2452536 & 2.12 $\\pm$ 0.11 & 173 $\\pm$ 2 & 1.28 $\\pm$ 0.11 & 10 $\\pm$ 2\\\\ & AAT & 15/08/03 & 2452867 & 2.35 $\\pm$ 0.11 & 179 $\\pm$ 1 & 1.61 $\\pm$ 0.11 & 11 $\\pm$ 1\\\\ & WHT & 30/09/04 & 2453279 & 3.40 $\\pm$ 0.15 & 174 $\\pm$ 2 & 1.93 $\\pm$ 0.15 & 11 $\\pm$ 2\\\\ \\hline \\end{tabular} \\caption{New polarimetric observations of HD~179821 and IRC~+10420 measured in the $R$ band. For the spectropolarimetric data (those taken at the AAT and WHT), the polarization was measured in the continuum region close to H$\\alpha$, while the final columns are the polarizations at the line centers. The H$\\alpha$ line-center polarization of HD~179821 was not calculated as no line effect was observed, and therefore this measurement would be equal to the continuum polarization. The systematic error of the AAT and WHT spectropolarimetric data is estimated to be of order 0.10\\% and 0.15\\%, respectively. } \\label{T:SUMDATA} \\end {table*} \\begin {table*} \\centering \\begin {tabular}{llrrrrrrrr} \\hline Julian Date & Telescope & $B-V$& $V$& $V-R$& $R-I$ & $J$ & $H$ & $K$ \\\\ % \\hline 2450268 & CST & & & & & 5.62& & \\\\ % 2450303 & CST & & & & & 5.36& 4.40& 3.45 \\\\ % 2450643 & CST & & & & & 5.40& 4.44& 3.50 \\\\ % 2450690 & CST & & & & & 5.40& 4.45& 3.49 \\\\ % 2450707 & CST & & & & & 5.37& 4.43& 3.38 \\\\ % 2450950 & NOT & 2.76 & 11.06& 2.40& 1.70& & & \\\\ % 2450985 & CST & & & & & 5.38& 4.44& 3.48 \\\\ % 2451037 & TSAO & 2.58 & 11.12& 2.42& 1.60& & & \\\\ % 2451039 & TSAO & 2.67 & 11.11& 2.41& 1.58& & & \\\\ % 2451042 & TSAO & 2.58 & 11.13& 2.42& 1.54& & & \\\\ % 2451043 & TSAO & 2.73 & 11.06& 2.49& 1.61& & & \\\\ % 2451047 & TSAO & 2.51 & 11.01& 2.41& 1.58& 5.63& 4.55& 3.73 \\\\ % 2451048 & TSAO & 2.75 & 11.03& 2.47& 1.60& 5.43& 4.57& 3.58 \\\\ % 2451050 & TSAO & 2.67 & 11.01& 2.46& 1.66& & & \\\\ % 2451052 & TSAO & 2.75 & 11.03& 2.47& 1.59& 5.35& 4.44& 3.53 \\\\ % 2451057 & TSAO & 2.76 & 10.96& 2.47& 1.64& & & \\\\ % 2451063 & TSAO & 2.60 & 10.95& 2.43& 1.59& & & \\\\ % 2451075 & TSAO & 2.70 & 11.16& 2.52& 1.60& & & \\\\ % 2451082 & TSAO & 2.66 & 11.09& 2.47& 1.59& & & \\\\ % 2451083 & TSAO & 2.72 & 11.05& 2.48& 1.60& & & \\\\ % 2451087 & CST & & & & & 5.24& 4.29& 3.33 \\\\ % 2451099 & TSAO & 2.62 & 11.16& 2.56& 1.66& & & \\\\ % 2451100 & TSAO & 2.71 & 11.16& 2.54& 1.66& & & \\\\ % 2451103 & TSAO & 2.76 & 11.13& 2.51& 1.54& & & \\\\ % 2451104 & TSAO & 2.72 & 11.03& 2.48& 1.59& & & \\\\ % 2451147 & CST & & & & & 5.35& 4.42& 3.46 \\\\ % 2451245 & TSAO & & & & & 5.32& 4.47& 3.56 \\\\ % 2451292 & CST & & & & & 5.40& 4.45& 3.57 \\\\ % \\hline \\end{tabular} \\caption{ Photometry of IRC~+10420. The data come from the Carlos S\\'{a}nchez Telescope (CST), the Nordic Optical Telescope (NOT) and the Tien-Shan Astronomical Observatory (TSAO), see text for details. Typically, the photometric errors are of order 0.01-0.03 magnitude.} \\label{T:PHOTDATA} \\end {table*} ", "conclusions": "We have presented spectropolarimetry and long term photo-polarimetry for two post-Red Supergiants. A strong depolarization across the H$\\alpha$ emission line is found for IRC~+10420, suggesting an electron-scattering region that is not circularly symmetric, confirming the results of \\cite{Davies_etal:2007}, who found such evidence from their integral field spectroscopy of the object. The time evolution indicates that the source has increased in temperature until at least 1995, after which the {\\it J} band photometry indicates this may have levelled off. If the temperature increase of the object has indeed halted, an increase in mass loss or mass infall rate can be responsible for the observed H$\\alpha$ emission and polarization. HD~179821, a cooler object, has less H$\\alpha$ emission, and no depolarization across the H$\\alpha$ line could be detected. The photometry implies that this evolved object is also undergoing a change in temperature. Hitherto, temperature determinations of the object were inconclusive, but if the star is a G supergiant, then the observed change in photometry suggests it has become earlier by one or two subclasses. Strong changes at the 5 per cent level in polarization over the past 15 years have been detected. During the same time, the optical photometry has only varied by at most 0.2 magnitudes. The most obvious explanations for this observation such as changes in either electron or dust scattering can be ruled out. A complication is that the polarization is not correlated with any other observable. The possibility that the star undergoes irregular, asymmetric, mass ejections is discussed. However, there is little evidence for the occurrence of such events. Instead, it is proposed that the star itself is asymmetric and that its anisotropic radiation, which is scattered off the circumstellar dust results in the net polarization observed. \\subsection*" }, "0801/0801.4083_arXiv.txt": { "abstract": "Using the OGLE catalogue of eclipsing binaries, 15 contact binaries were identified towards the SMC and the LMC at vertical distances from the Galactic plane between 300 pc and 10 kpc. Based on the luminosity function calculated for these contact binaries, we estimated a frequency of occurrence relative to Main Sequence stars in the thick disk at roughly $\\frac{1}{600}$. This estimate suffers from the small number statistics, but is consistent with the value previously found for the solar neighbourhood. ", "introduction": "During the past few decades, several attempts have been made to determine the local spatial density of contact binaries, with very diversified results. \\citet{ru02} (where full references are given and differences discussed) estimated their local spatial density at $(1.02\\pm0.24) \\times 10^{-5} \\ \\mathrm{pc^{-3}}$. Later, on the basis of the All Sky Automated Survey (ASAS), a relative frequency of occurrence (RFO) of one contact binary among about 500 solar type stars was derived \\citep{ru06}, which is in full agreement with the mentioned spatial density. The RFO for contact binaries at high galactic latitudes, however, has been entirely unknown. Here we present an estimate for the spatial occurrence of contact binary systems relative to Main Sequence (MS) stars in the thick disk of our Galaxy, far from the Galactic plane. In particular, the luminosity function for contact binaries in two conical volumes towards the Small Magellanic Cloud (SMC) and the Large Magellanic Cloud (LMC) covering parts of the thick disk and the halo is determined. The results are then compared with the luminosity function for MS stars in the same region of the sky. The term ``contact binaries'' is here used as a synonym for W~UMa-type eclipsing binaries with orbital periods in a range of 0.22 -- 1 days. In this paper, for reasons to be explained below, we limit ourselves to a subset with periods $<0.45$ days. \\section[Identification of contact binaries]% {Identification of short period contact binaries in the OGLE-catalogue of eclipsing binaries} The Optical Gravitational Lensing Experiment (OGLE) was intended to detect dark matter in the Milky Way Galaxy using the microlensing technique, with the Magellanic Clouds and the Galactic Bulge being the main targets of the survey \\citep{usk97}. As a by-product, OGLE provides high quality, long-term photometry that can be used to analyse eclipsing binary stars. Two online catalogues \\citep{wy03,wy04} were used for this study. They contain data in the standard photometric $BVI$ system for 2850 and 1351 eclipsing binaries, and cover an area of 4.6 and 2.4 square degrees of the central parts of the LMC and the SMC respectively. The OGLE-II survey has a faint limit of roughly $I=20.5 \\ \\mathrm{mag}$ with a corresponding error of $0.3 \\ \\mathrm{mag}$; the error becomes smaller for decreasing magnitude, reaching a bright limit of the survey at about 13 mag \\citep{wy04}. The differentiation between contact binaries and other eclipsing binary types in the two OGLE catalogues was carried out by applying a contact binary criterion based on Fourier analysis of the light curves, which was introduced by \\citet{ru97}\\footnote{More detailed explanations for this and other techniques used in this paper can be found in \\citep{ru97} and \\citep{ru06}.}. By means of this criterion and using a visual inspection of the remaining light curves, we identified 10 and 5 contact binaries with orbital periods $P<0.45 \\ \\mathrm{d}$ towards the LMC and the SMC, respectively. The light curves are plotted in Figure \\ref{fig_multi_plot}. Furthermore, we used the online database from the MACHO (Massive Compact Halo Object) project \\citep{al97} to confirm the classification of the contact binaries in the OGLE catalogue. A direct comparison with the OGLE catalogue was only possible for the contact binaries towards the LMC, because there is no MACHO photometry available for most survey fields covering the SMC. The MACHO photometry of the 10 contact binaries towards the LMC is in good agreement with the results obtained from the OGLE survey\\footnote{As an exception, the OGLE photometry in $V$-band for the contact binary OGLE050905.22-693315.1 was found to have a gross error of roughly 3 magnitudes.}. The contact binaries in our sample have orbital periods in a range of 0.22 -- 0.45 days. The short period limit is a natural cut-off for contact binaries \\citep{ru07}, whereas the upper period limit was intentional: On one hand, we wanted to be sure of the contact binary classification and avoid semi-detached binaries, while on the other hand, we wanted to use the simple $M_V \\equiv M_V (\\log P)$ calibration and avoid problems with uncertain or missing colour indices. Because the final RFO estimate is done using absolute magnitude bins of the luminosity function, the period limits signify an intentional restriction to the low brightness end of the contact binary sequence. We will explain the details below. ", "conclusions": "We conclude that each of the two samples yields the relative frequency of one contact binary among about $600$ MS stars in the two conical volumes towards the SMC and the LMC at galactic latitudes of $-44\\degr$ and $-33\\degr$ respectively. These estimates have uncertainties of roughly $50 \\%$ due to the large uncertainties arising from small number statistics and to the still uncertain parameters of the spheroidal component of the Galaxy. This component has been explicitly accounted for in the previous section and is estimated to contribute to the total number of stars in the SMC and the LMC search volumes at levels of about 30\\% and 20\\%, respectively. The contribution varies with the $M_V$. Specifically, for the SMC, it changes from 38\\% for the $M_V$-bin with the largest distances to 17\\% for the $M_V$-bin centred at 6 mag; for the LMC, it changes from 29\\% to 10\\% in the same range of $M_V$. The results on the relative frequency of contact binary stars at large galacto-centric distances is consistent with estimates for the solar neighbourhood in previous work \\citep{ru06}. These are the very first data available on the contact binary distribution at high galactic latitudes. \\begin{figure} \\begin{center} \\includegraphics[width=84mm]{lf_lmc.eps} \\caption{\\label{fig-lf_lmc}Luminosity function towards the LMC. The continuous line represents the contact binary luminosity function, whereas the dashed line represents the MS luminosity function divided by a factor of 650.} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=84mm]{lf_smc.eps} \\caption{\\label{fig-lf_smc} Luminosity function towards the SMC. The same as in figure \\ref{fig-lf_lmc} except for the scaling factor of 600 instead of 650.} \\end{center} \\end{figure} \\begin{figure*} \\begin{center} \\includegraphics[width=176mm]{multi_plot.eps} \\caption{\\label{fig_multi_plot} The light curves of the 15 close contact binaries. The fitting was done according to the approach of \\citep{ru93} (i.e. using only the first 5 cosine and the first sine Fourier-coefficients).} \\end{center} \\end{figure*}" }, "0801/0801.4560_arXiv.txt": { "abstract": "Narrow Line Seyfert 1 galaxies (NLS1s) are generally considered peculiar objects among the broad class of Type 1 active galactic nuclei, due to the relatively small width of the broad lines, strong X--ray variability, soft X--ray continua, weak \\Oiii{}, and strong \\Feii{} line intensities. The mass \\mbh{} of the central massive black hole (MBH) is claimed to be lighter than expected from known MBH--host galaxy scaling relations, while the accretion rate onto the MBH larger than the average value appropriate to Seyfert 1 galaxies. In this {\\it Letter}, we show that NLS1 peculiar \\mbh{} and $L/L_{\\rm Edd}$ turn out to be fairly standard, provided that the broad line region is allowed to have a disc--like, rather than isotropic, geometry. Assuming that NLS1s are rather ``normal'' Seyfert 1 objects seen along the disc axis, we could estimate the typical inclination angles from the fraction of Seyfert 1 classified as NLS1s, and compute the geometrical factor relating the observed FWHM of broad lines to the virial mass of the MBH. We show that the geometrical factor can fully account for the ``black hole mass deficit\" observed in NLS1s, and that $L/L_{\\rm Edd}$ is (on average) comparable to the value of the more common broad line Seyfert 1 galaxies. ", "introduction": "Seyfert 1 galaxies (Sy1s) are often divided into two distinct classes, namely Broad Line Sy1s (BLS1s), whose H$\\beta$ line has FWHM~$\\gsim 2000$ km/s (hence, as standard Type 1 AGN), and Narrow Line Sy1s (NLS1s), with lower velocities (e.g., Goodrich, 1989). NLS1s are a minority, $\\simeq15$\\% of all the Sy1s, according to the optical spectroscopic classification of the SDSS general field (Williams Pogge \\& Mathur, 2002), the fraction depending on the AGN luminosity (with a peak at $M_{g'}\\sim -22$), and on the radio loudness (radio loud NLS1s account only for $\\sim7$\\% of the class, Komossa et al., 2006, but it is still debated if the NLS1s can be considered a peculiar radio-quiet sub-class among Sy1s, see e.g. Sulentic et al. 2007; Doi et al. 2007). NLS1s also show weak \\Oiii{} and strong \\Feii{} emission line (Osterbrock \\& Pogge, 1985), strong variability, and a softer than usual X--ray continuum (Boller Brandt \\& Fink, 1996; Grupe et al., 1999). Grupe \\& Mathur (2004) found that NLS1s have, on average, lower \\mbh{} than expected from \\mbh{}--host galaxy relations such as \\mbh{}--$\\sigma_*$ (see Tremaine et al., 2002, and references therein), while BLS1 \\mbh{} are in fairly good agreement to the same relation. The estimated low values of \\mbh{} lead to an average Eddington ratio $L/L_{\\rm Edd}$ for the NLS1 population which is almost an order of magnitude larger than the average value of BLS1s ($L/L_{\\rm Edd} \\simeq 1$ to be compared to $\\simeq 0.1$, Grupe, 2004). Further evidence of low \\mbh{} in NLS1s comes from the observed rapid X--ray variability (see., e.g., Green, McHardy \\& Lehto 1993, and Hayashida 2000). Such results were interpreted as indications of a peculiar role of NLS1s within the framework of the cosmic evolution of MBHs and of their hosts. In a MBH-galaxy co--evolution scenario, NLS1s are thought to be still on their way to reach the \\mbh{}-$\\sigma_*$ relation, i.e., their (comparatively) small MBHs are highly accreting in already formed bulges. Recently Botte et al. (2005) and Komossa and Xu (2007) came to the conclusion that NLS1s have indeed smaller masses and higher $L/L_{\\rm Edd}$ than BLS1, nevertheless they both do follow the $M-\\sigma_*$ relation for quiescent galaxies. The authors argued that the customarily used \\Oiii{} line is not a reliable surrogate for the stellar velocity dispersion $\\sigma_*$. The Grupe and Mathur's results and interpretation have been recently confirmed and supported by several other groups, see, e.g., Zhou et al. (2006) and Ryan et al. (2007). Ryan et al. (2007) pointed out that IR-based mass measurements might be unreliable because of the extra IR contribute from the circum-nuclear star-forming regions in NLS1s. Notwithstanding, they suggested that this contamination can not significantly affect their data, and thus is insufficient to account for the MBH mass deficit. In the aforementioned papers, \\mbh{} was computed as \\begin{equation}\\label{eq_virial} M_{\\rm BH} = \\frac{R_{\\rm BLR} v_{\\rm BLR}^2}{G}, \\end{equation} where $R_{\\rm BLR}$ is the broad line region (BLR) scale radius, and $v_{\\rm BLR}$ the typical velocity of BLR clouds. $R_{\\rm BLR}$ is found by means of the reverberation mapping technique (Blandford \\& McKee, 1982), or by exploiting statistical $R_{\\rm BLR}$--luminosity relations (see Kaspi et al., 2000, 2005 and 2007); $v_{\\rm BLR}$ can be inferred from the H$\\beta$ width as \\begin{equation}\\label{eq_def_f} v_{\\rm BLR} = f \\cdot {\\rm FWHM}, \\end{equation} where the FWHM refers only to the broad component of the line, and $f$ is a fudge factor which depends upon the assumed BLR model. For an isotropic velocity distribution, as generally assumed, $f=\\sqrt{3}/2$. Labita et al. (2006) and Decarli et al. (in preparation) found that in QSOs an isotropic BLR fails to reproduce the observed line widths and shapes, and a disc model should be preferred. A disc--like geometry for the BLR has been proposed by several authors in the past (e.g., Wills \\& Browne 1986; Vestergaard, Wilkes \\& Barthel 2000; Bian \\& Zhao, 2004). In this picture, the observed small FWHM of NLS1 broad lines are ascribed to a small viewing angle with respect to the disc axis, and no evolutionary difference is invoked whatsoever. In this {\\it Letter}, we adopt the disc--like model for the BLR of Seyfert galaxies. We use the observed frequency of NLS1s to estimate their typical viewing angle, and then compute the appropriate geometrical factor $f$. Using eq.~\\ref{eq_virial}, we will show that the new estimates of \\mbh{} for NLS1s nicely agree with the standard \\mbh{}--$\\sigma_*$ relation. In turn, the accretion rate of the class is found to be similar to that of BLS1s. ", "conclusions": "In this {\\it Letter}, we have assessed the claimed peculiarity of Narrow Line Seyfert 1 galaxies within the framework of cosmic evolution of massive black holes, and their host bulges. Indeed, the optical properties of NLS1s, their X--ray fast variability and the faintness of their bulges can be accounted for if one admits lower black hole masses and higher accretion rates (in Eddington units) than standard Broad Line Seyfert 1 galaxies (BLS1s), placing NLS1s in an early evolutionary stage (Grupe \\& Mathur, 2004; Grupe, 2004; Botte et al., 2004; Zhou et al., 2006; Ryan et al., 2007). If this is true, by observing local NLS1s we can have hints of the infancy of the ubiquitous population of super--massive black holes. We have explored an alternative explanation to the narrowness of H$\\beta$ lines in NLS1s, namely, pole--on orientation of a disc--like broad line region. If BLS1s and NLS1s differ only by the observation angle of the BLR disc, the frequency of NLS1s among the Sy1 class gives the limiting viewing angle of NLS1s. Then, assuming $H/R\\lsim 0.1$ for the disc, we computed corrected geometrical factors linking the observed FWHM to \\mbh{}, and found $f_{\\rm NLS1}\\gsim 2$ and $f_{\\rm BLS1}\\simeq 1$, in agreement with recent estimates given by Labita et al. (2006). The idea of a disc--like BLR is not new (e.g., Wills \\& Browne 1986; Vestergaard, Wilkes \\& Barthel 2000; Bian \\& Zhao, 2004), but for the first time, by re-calculating masses and Eddington ratios for a sample of Sy1s, we found that mass and luminosity functions are similar for NLS1s and BLS1s. In a sense, we can say that all Sy1s are normal, but some are more ``normal'' than others. We note that, though NLS1s seem to lie in the same region of the \\mbh--$\\sigma_*$ plane, the adopted $\\sigma_*$ values can be largely over--estimated (Komossa \\& Xu, 2007; Mullaney \\& Ward, 2007), and then firm conclusions on the \\mbh--$\\sigma_*$ issue can not be drawn at this stage. Can a simple orientation model, as the one we adopted here, explain the unique observed properties of NLS1s? NLS1s differ from standard Sy1s not just in the width of optical lines, but, more noticeably, in what are their X--ray properties, both spectral and temporal. The X--ray emission of NLS1s has been studied and discussed in great details by, among others, Boller et al. (1996), using {\\it ROSAT} data, and by Brandt, Mathur \\& Elvis (1997) using {\\it ASCA} data. NLS1s have generally both soft and hard X--ray spectra which are steeper than normal Sy1s, and show large amplitude, rapid variability. Boller et al. (1996) showed how different models, invoking extreme values of one or more of the followings: pole--on orientation, black hole mass, accretion rate, warm absorption, BLR thickness, all explain some aspects of the complex NLS1 soft X--ray phenomenology, but, still, all appear to have drawbacks. If pole--on orientation has to be the main cause of the uniqueness of the X--ray features of NLS1s, then a necessary condition is that the hard power--law emission is not intrinsically isotropic, e.g., a thermal extended corona (as in Haardt \\& Maraschi, 1991; 1993) is not a viable option. Models in which the X--rays of type I radio quiet AGNs are funneled or beamed have been proposed by several authors (e.g., Madau, 1988; Henri \\& Petrucci, 1997; Malzac et al., 1998; Ghisellini, Haardt \\& Matt, 2004). For example, Ghisellini et al. (2004) showed that an aborted jet model, in which X--rays are produced by dissipation of kinetic energy of colliding blobs launched along the MBH rotation axis, can explain the steep and highly variable X--ray power law. The model, in its existing formulation, does not allow clear predictions of spectral and temporal features other than in the X-rays. To assess its relevance for NLS1s would require a much more detailed modeling. In particular, the peculiar \\Feii{} and \\Oiii{} properties must be accounted for. The statistics of radio-loud NLS1s is low. In several works the existence of differences in the radio properties between NLS1s and BLS1s has been discussed (see, e.g. Komossa et al. 2006; Zhou et al. 2006; Sulentic et al. 2007; Doi et al. 2007). Doi et al. (2007) suggested that $\\sim$ 50 \\% of radio-loud NLS1s are likely associated with jets with high brightness temperatures, requiring Doppler boosting. This interpretation supports the pole--on orientation model (for a different point of view see Komossa et al. 2006). Our results, if confirmed, indicate that a population of accreting, undermassive MBHs (with respect to the \\mbh{}--$\\sigma_*$ relation) has to be found yet. This may suggest that the \\mbh{}--$\\sigma_*$ relation was established long ago, during the MBH accretion episodes following the first major mergers of the host galaxies. Moreover, Komossa \\& Xu (2007) found that NLS1s do follow the \\mbh{}--$\\sigma_*$ relation of non--active galaxies, but still they have smaller \\mbh{} and larger $L/L_{\\rm Edd}$ than BLS1s. If this is the case, then $\\sigma_*$ of the host bulges of NLS1 needs to evolve accordingly in order to preserve the \\mbh{}--$\\sigma_*$ relation, or, alternatively, NLS1s are the low mass extension of BLS1s, and the NLS1 high $L/L_{\\rm Edd}$ is a short--lived phenomenon. We note here that the interpretation of Komossa \\& Xu (2007), as well as the one of Grupe \\& Mathur (2004), implies that \\mbh{} and $L/L_{\\rm Edd}$, in principle independent quantities, somehow conspire to produce comparable luminosities as observed in NLS1s and BLS1s. Applying our correction to the \\mbh{} as well as the one to the $\\sigma_*$ proposed by Komossa \\& Xu (2007), the NLS1s would be even off--setted towards higher masses with respect to the \\mbh{}--$\\sigma_*$ relation. There are however two possible problems with the pole--on orientation model. First, according to the orientation model, the polarization properties of broad emission lines should depend on the inclination angle, in the sense that nearly pole--on Sy1s should not be polarized. However, Smith et al. (2004) found polarized broad lines in few NLS1s, and traces of broad H$\\alpha$ polarization were also found by Goodrich (1989) in 6 out of 17 NLS1s. A second issue is discussed by Punsly (2007), who finds larger line broadening in face--on quasars, possibly due to large isotropic gas velocities or winds. In conclusion, we found that orientation effects can account for the different optical properties of NLS1s compared to the more common BLS1s. The model is particularly appealing, as it naturally sets masses and accretion rates of NLS1 to fairly standard values. To validate this interpretation, orientation must be able to explain the extreme X--ray properties of NLS1. Jetted models for radio quiet AGNs may be promising in this, and we urge a detailed, critical comparison of such models with the bulk of NLS1 data." }, "0801/0801.1035_arXiv.txt": { "abstract": "{In this paper we investigate the radio-MIR correlation at very low flux densities using extremely deep 1.4\\,GHz sub-arcsecond angular resolution MERLIN$+$VLA observations of a 8\\farms5$\\times$8\\farms5 field centred upon the Hubble Deep Field North, in conjunction with \\spitzer\\ 24\\,\\mum\\ data. From these results the MIR-radio correlation is extended to the very faint ($\\sim$microJy) radio source population. Tentatively we detect a small deviation from the correlation at the faintest IR flux densities. We suggest that this small observed change in the gradient of the correlation is the result of a suppression of the MIR emission in faint star-forming galaxies. This deviation potentially has significant implications for using either the MIR or non-thermal radio emission as a star-formation tracer of very low luminosity galaxies.} ", "introduction": "Since 1970s and 1980s studies of the radio and far-infrared (FIR) properties of galaxies have shown there to be a tight correlation between their emission in these two observing bands which extends over several orders of magnitude in luminosity \\citep{vanderkruit73,condon82}. The advent of the {\\it Infrared Astronomical Satellite} (IRAS) All Sky Survey in 1983 \\citep{neugebauer84,soifer87} enabled much larger systematic samples of galaxies to be studied at infrared wavelengths and hence further demonstrated the consistency and tightness of this correlation, albeit for relatively nearby sources \\citep{helou85,dejong85,condon86,condon91,yun01}. Following {\\it IRAS}, deep observations using the {\\it Infrared Space Observatory} ({\\it ISO}) allowed fainter and higher redshift galaxies to be observed at mid-infrared (MIR) wavelengths. These {\\it ISO} observations showed that the MIR emission from galaxies is loosely correlated with radio emission across a wide range of redshifts, tentatively extending to z$\\sim$4 \\citep{cohen00,elbaz02,garrett02}. More recently the launch of the \\spitzer\\ Space Telescope in August 2003 has greatly increased the sensitivity of MIR observations and hence our ability to study the MIR-radio correlation. Early results, such as from the \\spitzer\\ First Look Survey (FLS), have confirmed that the MIR-radio correlation holds for relatively bright star-forming galaxies (S$_{20\\,{\\rm cm}}>115$\\,$\\mu$Jy) out to at least a redshift of 1 \\citep{appleton04}. Radio and infrared emission from galaxies in both the nearby and distant Universe is thought to arise from processes related to star-formation, hence resulting in the correlation between these two observing bands. The infrared emission is produced from dust heated by photons from young stars and the radio emission predominately arises from synchrotron radiation produced by the acceleration of charged particles from supernovae explosions. It has however recently been suggested that at low flux density and luminosities there may be some deviation from the tight well-known radio-IR correlation seen for brighter galaxies \\citep{bell03,boyle07}. \\begin{center} \\begin{figure*} \\setlength{\\unitlength}{.5in} \\begin{picture}(16,5) \\put(-0.5, 0){\\special{psfile=Beswick-fig1a.ps hscale=90 vscale=90}} \\put(6.5, 0){\\special{psfile=Beswick-fig1b.ps hscale=90 vscale=90}} \\end{picture} \\vskip -0.5cm \\caption{{\\it Left-hand panel:} Radio 1.4\\,GHz versus the MIR 24\\,\\mum\\ flux density of all 377 individual sources (small triangle), and median radio flux density logarithmically binned by their 24\\,\\mum\\ flux density (filled circles) within the 8\\farms5$\\times$8\\farms5 field. {\\it Right-hand panel:} Control plot of 1.4\\,GHz radio flux densities plotted against the 24\\,\\mum\\ source flux densities of the sample. The radio flux densities of the control sample have derived in exactly the same manner as the flux densities plotted in the left-hand, panel but have been measured at randomly assigned sky positions in the 8\\farms5$\\times$8\\farms5 radio image where no known sources at any wavelength are located.} \\label{F20vsF24} \\end{figure*} \\end{center} \\cite{bell03} argue that while the IR emission from luminous galaxies will trace the majority of the star-formation in these sources, in low luminosity galaxies the IR emission will be less luminous than expected considering the rate of star-formation within the source (i.e. the IR emission will not fully trace the star-formation). In this scenario the reduced efficiency of IR production relative to the source star-formation rate (SFR) would be the result of inherently lower dust opacities in lower luminosity sources and consequently less efficient reprocessing of UV photons from hot young stars into IR emission. The simple consequence of this is that at lower luminosities the near linear radio-IR correlation L$_{\\rm radio}\\propto$L$_{\\rm IR}^{\\!\\gamma}$, with $\\gamma >1$ \\citep[e.g][]{cox88,price92} will be deviated from. {\\it Of course such an assertion is dependent upon the radio emission providing a reliable tracer of star-formation at low luminosities which may be equally invalid.} Recently \\cite{boyle07} have presented a statistical analysis of Australia Telescope Compact Array (ATCA) 20\\,cm observations of the 24\\,\\mum\\ sources within two regions (the {\\it Chandra} Deep Field South (CDFS) and the European Large Area {\\it ISO} Survey S1 (ELAIS)) of the \\spitzer\\ Wide Field Survey (SWIRE). In this work \\cite{boyle07} have co-added sensitive (rms$\\sim$30\\,$\\mu$Jy) radio data at the locations of several thousand 24\\,\\mum\\ sources. Using this method they have statistically detected the microJy radio counterparts of faint 24\\,\\mum\\ sources. At low flux densities (S$_{\\rm 24\\, \\mu m }=100\\,\\mu$Jy) they confirm the IR-radio correlation but find it to have a lower coefficient (S$_{\\rm 1.4\\,GHz}$\\,=\\,0.039\\,S$_{\\rm 24\\,\\mu m}$) than had previously been reported at higher flux densities. This coefficient is significantly different from results previously derived from detections of individual objects (e.g. \\citealt{appleton04}) and is speculated by \\cite{boyle07} to be the result of a change in the slope of the radio-IR correlation at low flux densities. In this paper, we utilise very deep, high resolution 20\\,cm observations of the Hubble Deep Field North and surrounding area made using MERLIN and the VLA \\citep{muxlow05} in combination with publicly available 24\\,\\mum\\ \\spitzer\\ source catalogues from GOODS to study the MIR-Radio correlation for microjansky radio sources. This study extends the flux density limits of the radio-IR correlation by more than an order of magnitude for individual sources and overlaps the flux density regime studied using statistical stacking methods by \\cite{boyle07}. Additionally we employ statistical stacking methods, similar to those used by \\cite{boyle07}, to extend the correlation further to still lower flux densities. We adopt H$_0=$75\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_{\\rm m}=0.3$ and $\\Omega_{\\rm \\Lambda}$=0.7 throughout this paper. ", "conclusions": "Using one of the deepest high-resolution 1.4\\,GHz observations made to date, in conjunction with deep 24\\,\\mum\\ \\spitzer\\ source catalogues from GOODS, we have investigated the microJy radio counterparts of faint MIR sources. These observations confirm that the microJy radio source population follow the MIR-radio correlation and extend this correlation by several orders of magnitude to very low flux densities and luminosities, and out to moderate redshifts. This extension of the MIR-radio correlation confirms that the majority of these extremely faint radio and 24\\,\\mum\\ sources are predominantly powered by star-formation with little AGN contamination. Statistically stacking the radio emission from many tens of faint 24\\,\\mum\\ sources has been used to characterise the size and nature of the radio emission from very faint IR galaxies well below the nominal radio sensitivity of these data. Using these methods the MIR-radio correlation has been further extended and a tentative deviation in this correlation at very low 24\\,\\mum\\ flux densities has been identified. \\subsection*" }, "0801/0801.0343_arXiv.txt": { "abstract": "Galaxy disks evolve through angular momentum transfers between sub-components, like gas, stars, or dark matter halos, through non axi-symmetric instabilities. The speed of this evolution is boosted in presence of a large fraction of cold and dissipative gas component. When the visible matter dominates over the whole disk, angular momentum is exchanged between gas and stars only. The gas is driven towards the center by bars, stalled transiently in resonance rings, and driven further by embedded bars, which it contributes to destroy. From a small-scale molecular torus, the gas can then inflow from viscous torques, dynamical friction, or $m=1$ perturbations. In the weakened bar phases, multiple-speed spiral patterns can develop and help the galaxy to accrete external gas flowing from cosmic filaments. The various phases of secular evolution are illustrated by numerical simulations. ", "introduction": "\\subsection{Bar formation and evolution} Non-axisymmetries and in particular bars are the motor of secular evolution, in transferring the angular momentum in galaxies. In the 80' and 90', numerical simulations established how bars formed, in pure stellar disks, without any dark matter haloes, or embedded in rigid halo components: the angular momentum was exchanged within the disk, the outer parts gaining the momentum from the inner parts. This exchange occured mainly at bar formation, the bar pattern speed slowing down slightly, while the stellar orbits in the bar were more and more elongated, corresponding to a lower precessing rate. Once formed, the bar was then robust (e.g. Sellwood 1981). Already, some dynamical heating was observed to produce some feedback in the bar strength evolution: more unstable stellar disks develop a bar sooner, but then end up with a weaker bar than more stable disks (Combes et al 1990). \\subsection{Bars with gas disks} The introduction of gas in disks completely changes the stability, due to dissipation. The angular momentum can be exchanged between gas and stars, and from inner to outer parts, again with a negligible halo component. While gas flows towards the center, the bar is destroyed and gives rise to a triaxial bulge (Friedli \\& Benz 1993). All this evolution can be traced back to the bar gravity torques acting on the gas. The torques are proportional to the phase shift between the gas and the potential wells due mainly to the stars in the bar. Dissipation produces these phase shifts, which are conspicuous through the characteristic leading dust lanes in barred galaxies. The gas then loses angular momentum and mass concentrates towards the center. The amplitude of the phenomenon has been quantified from observations of barred galaxies. The gravity torques can be computed from the stellar potential deduced from the red image of the galaxy, and the effective angular momentum exchange computed from the gas distribution, obtained through H$\\alpha$, HI or CO emission images. In strongly barred galaxies, the gas inside corotation inflows to the center in one or two rotations. Even in weakly barred galaxies, the gravity torques are efficiently producing gas flows, and fueling the nucleus, as for instance in NGC 6574 (Lindt-Krieg et al 2007) or NGC 3147 (Casasola et al 2008). \\subsection{Formation of rings} The secular evolution as described above, the gas inflow driven by bars, finds its confirmation in the frequent observation of resonant rings, where gas is piling up, and form stars in bright knots. The gravity torques change sign at each resonance, and cancel in the rings where the gas disctribution is symmetrically distributed with respect to the stars. Galaxies often possess multiple-rings, corresponding to the various Lindblad resonances, ILR, UHR, OLR (Buta \\& Combes 1996). With only one bar, the gas stops its inflow at the ILR, in the nuclear ring. The decoupling of a nuclear bar inside the nuclear ring triggers the AGN fueling (cf NGC 2782, Fig 1). \\begin{figure} \\begin{center} \\includegraphics[angle=-90,width=12cm]{combes-fig1.ps} \\end{center} \\caption{ Gray scale plot of the stellar component ({\\it left}), the gas component ({\\it middle}), and an expanded version of the gas distribution ({\\it right}), in a self-consistent simulation of a double bar formation, meant to reproduce the galaxy NGC 2782 (the axes are in kpc). It can be seen that the gas, which was predominantly in the more external ring, corresponding to the ILR of the primary bar, at T=690 Myr, is falling progressively inward, and is found inside the ILR at T=750 Myr. From Hunt et al (2008).} \\label{n2782} \\end{figure} \\vspace{-0.3cm} ", "conclusions": "\\vspace{-0.2cm} Secular evolution is important for bars, it controls their strength, their pattern speed, or their vertical thickness through peanut formation. The dominant angular momentum transfer from the disk occurs with the DM halo or with the outer disk, according to the halo to disk mass ratio, or between the stellar and gaseous components, according to the dissipation character of the gas. Bars are weakened by central mass concentrations and/or gas inflows, driven by bars themselves, implying self-regulation. The gas fraction able to destroy bars depends on the dark matter halo to disk mass ratio. Lopsidedness ($m=1$ global mode) can also weaken the bar. Bars dissolve into lenses, especially in early-type galaxies. \\vspace{-0.3cm}" }, "0801/0801.2164_arXiv.txt": { "abstract": "\\noindent Recent data support the idea that the filaments observed in H$\\alpha$ emission near the centres of some galaxy clusters were shaped by bulk flows within their intracluster media. We present numerical simulations of evaporated clump material interacting with impinging winds to investigate this possibility. In each simulation, a clump falls due to gravity while the drag of a wind retards the fall of evaporated material leading to elongation of the tail. However, we find that long filaments can only form if the outflowing wind velocity is sufficiently large, $\\sim 10^{8}\\,{\\rm cm\\,s^{-1}}$. Otherwise, the tail material sinks almost as quickly as the cloud. For reasonable values of parameters, the morphological structure of a tail is qualitatively similar to those observed in clusters. Under certain conditions, the kinematics of the tail resemble those reported in Hatch et al.(2006). A comparison of the observations with the numerical results indicates that the filaments are likely to be a few tens of Myrs old. We also present arguments which suggest that the momentum transfer, from an outflowing wind, in the formation of these filaments is probably significant. As a result, tail formation could play a role in dissipating some of the energy injected by a central AGN close to the cluster centre where it is needed most. The trapping of energy by the cold gas may provide an additional feedback mechanism that helps to regulate the heating of the central regions of galaxy clusters and couple the AGN to the ICM. ", "introduction": "Optical emission-line nebulae commonly surround massive galaxies in the centres of X-ray bright, cool cluster cores \\citep[][]{crawf}. The origin of the H$\\alpha$ filaments has been attributed to a variety of processes including: condensation from an intracluster medium (ICM) evolving as a cooling flow \\citep[][]{fab84, heckman, don91}; accretion of clouds captured in galaxy mergers \\citep[][]{braine}; expulsion from the central galaxy \\citep[][]{burb}. NGC 1275 at the centre of the Perseus cluster contains the best studied example of such a nebula \\citep[][]{hatch}. Its filaments are typically 50-100 pc thick and up to 30 kpc long, and the majority of them are radial. \\cite{hatch} presented kinematic data that rules out dynamical models of purely infalling filaments. The observed kinematic properties provide strong evidence that the filaments are not in gravitational free fall, because, if they were, their velocities would rise sharply towards the centre of the nebula \\citep[][]{heckman}. The most conclusive evidence lies in the velocity structures of the northern and northwestern filaments. The lower half of the northern filament is redshifted with respect to the galaxy, whilst the upper section is blueshifted. Thus, the upper part of the filament is flowing away from the galaxy while the lower part is flowing into the galaxy. The filaments may follow streamlines of the flow of more tenuous material around them. In the Perseus cluster, some filaments appear to have been shaped by the wakes of the buoyant bubbles. Given that the data suggest that, at least, some parts of the filaments are outflowing, their origin may lie within the galaxy. NGC 1275 contains a large reservoir of cold molecular gas that could fuel them. The filaments have morphologies similar to those of laminar jets, and their relatively smooth structures may place constraints on turbulent motions in the ICM. Alternatively, an ordered, amplified magnetic field trailing behind a buoyant bubble interacting with filaments may prevent the destruction of the filaments by turbulent motions \\citep[][]{rusz07}. The optical nebula around NGC 1275 emits $4.1 \\times 10^{42}{\\rm erg\\,s^{-1}}$ in H$\\alpha$ and N[II]. The nature of the power source remains unclear. Various excitation mechanisms for these lines have been proposed. For instance, although ionisation by the central AGN may be important for the inner regions, it cannot be the dominant source of power for the extended nebula because the H$\\alpha$ intensity does not decrease with distance from the nucleus \\citep[][]{john88}. The nebula may be excited by stellar UV, but there is no spatial correlation between the filaments and the stellar clusters. Excitation by X-rays from the ICM seems unlikely, as the ICM may be as much as a hundred times less luminous in UV than in the X-rays \\citep[][]{fab03b}. Heat conduction from the ICM to the colder filaments has also been proposed \\citep[][]{don00}, but it might also lead to the evaporation of the filaments on too short a timescale \\citep[][]{nb04}. Shocks and turbulent mixing layers might play a role \\citep[][]{crawf92}. Cosmic rays, preferentially diffusing along the magnetic field lines trailing behind rising bubbles, could possibly drive the excitation in those filaments that are located in bubble wakes \\citep[][]{rusz07}. The same magnetic fields lines could also prevent the filaments from evaporating due to thermal conduction. The same mechanisms that power the emission could also have a profound effect on the morphology of the emitting region. It is impossible to take all of these processes into account. Consequently, we will investigate the effect of gravity and an impinging wind that strips material from a cloud on the morphology and kinematics of the resulting filament. The main aim of the work reported here is the calculation of the density and velocity structures of material evaporated from clumps embedded in winds from the central galaxies of galaxy clusters and the comparison of model results with observations. Section 2 contains some preliminary considerations, while the model assumptions and numerical approach are treated in section 3. The simulation results appear in section 4. In section 5 we investigate the development of tails in realistic environments by including the appropriate density and gravity for a galaxy cluster. Section 6 concerns momentum tranfer between clumps and the winds surrounding them. Section 7 discusses the possible trends in optical emission between clusters, and section 8 is a summary. ", "conclusions": "This work is intended mainly as a preliminary study into the processes that shape filamentary structures in the ICM, and not an exhaustive study. It is likely that there are many other processes such as magnetic fields, thermal conduction and viscosity that all play a role in the dynamics that produce the filamentary kinematics and morphology. However, we have concentrated on the basic hydrodynamic system and have found several interesting mechanisms that may go some way to explaining the properties of the observed filaments. The first ensemble of simulations show that the long, relatively straight filaments at the centre of the Perseus cluster may be formed by the interaction of a wind with clumps of cold material. Flows of low Mach number winds generate structure in the filaments which may be comparable with observations. These results also suggest that the amplitude of the velocity fluctuations along the length of the filament grow with increasing density contrast between the cold material and the ambient wind. The morphology of the observed filaments suggests that the density contrast is large ($\\sim 10^{4}$). Interestingly, the fluctuations evolve with time providing a possible diagnostic tool for estimating the ages of filaments from observations. in this set of simulations, the best comparison with observations suggests filament ages of $\\sim 40$ Myrs in the Perseus cluster. In a more realistic environment with spatially varying gravity, based on the Perseus cluster, filaments only form if the wind velocity is sufficiently large. If there is no significant wind, then the optical emission may be more amorphous, or more likely, filamentary on much shorter length scales. The morphology and kinematics suggests that a density contrast of $\\eta \\sim 10^{3}$ may be more compatible with the observations, while the best agreement in the kinematic data occurs for $\\eta \\sim 10^{4}$. As a compromise it is possible that the real conditions lay somewhere between these two extremes. Alternatively, other physical processes could be important. It should also be noted that these were 2-d simulations, and a full 3-d study would really be required to find the model conditions that best match reality. Regardless of this, the results do point to a relatively narrow range of parameters providing possible information about the state of the ICM, and the age of the filaments. These results suggest that the cold material is of the order of $10^{4}$ times denser than the ambient ICM, there is an outflowing wind of velocity of $\\sim 10^{3}\\,{\\rm km\\,s^{-1}}$, and that the filaments are approximately a few 10's of Myrs old. The masses, and densities, of the filaments created by these simulations seem to be low compared with values presented in the previous section. The simulations produce filaments of $10^{6}-10^{7}\\,{\\rm M_\\odot}$, while the crude estimates above suggested that $10^{8}\\,{\\rm M_\\odot}$ filaments might be more likely. It should be noted though, that the quantity of cold gas, in the simulated tail, is more than capable of producing the observed optical luminosity. The simulations also show a lot of cold, dense gas in the vicinity of the cloud itself. We also highlight the possibility that momentum transfer between the ICM and cold clouds could be a dynamically important process in galaxy clusters. This process could effectively couple the energy and momentum of an outflowing wind to the cold material, thereby dissipating some of the energy injected by a central AGN. The presence of cold gas may therefore provide an additional feedback mechanism that traps energy in the central regions of clusters where it is most needed." }, "0801/0801.2022_arXiv.txt": { "abstract": "The understanding and modeling of the structure and evolution of stars is based on statistical physics as well as on hydrodynamics. Today, a precise identification and proper description of the physical processes at work in stellar interiors are still lacking (one key point being that of transport processes) while the comparison of real stars to model predictions, which implies conversions from the theoretical space to the observational one, suffers from uncertainties in model atmospheres. That results in uncertainties on the prediction of stellar properties needed for galactic studies or cosmology (as stellar ages and masses). In the next decade, progress is expected from the theoretical, experimental and observational sides. I illustrate some of the problems we are faced with when modeling stars and the possible tracks towards their solutions. I discuss how future observational ground-based or spatial programs (in particular those dedicated to micro-arc-second astrometry, asteroseismology and interferometry) will provide precise determinations of the stellar parameters and contribute to a better knowledge of stellar interiors and atmospheres in a wide range of stellar masses, chemical compositions and evolution stages. ", "introduction": "Major goals of stellar structure and evolution studies are (i) to characterize and describe the physics of matter in the extreme conditions encountered in stars and (ii) to determine stellar properties (like age and mass) that trace the history and evolution of galaxies and constrain cosmological models. To achieve these goals, we rely on numerical stellar models based on input physics that integrate the results of recent theoretical studies, numerical simulations and laboratory experiments. The models inputs and outputs are chosen and/or validated by comparison with accurate astronomical observations. Numerical 2 and 3{\\small D} hydrodynamical simulations of limited regions of stellar interiors and atmospheres are now under reach of computers. They provide valuable constraints and data for current standard (1{\\small D}) stellar models: abundances, convection, rotationally induced instabilities and mixing, magnetic fields, etc. \\citep[see e.g.][for reviews]{2005ARA&A..43..481A,2007arXiv0708.1499T,2007IAUS..239..517Z}. In parallel, the physics of stellar plasmas is studied in the laboratory with (i) fluid experiments \\citep[study of turbulence in rotating, magnetic fluids, etc., see e.g.][]{1999A&A...347..734R}, (ii) particle accelerators (nuclear reaction cross sections, etc.) and, (iii) the so-called high energy-density facilities (based on high power lasers or z-pinches) which aim at exploring the high temperature and high density regimes found in stars, brown dwarfs and giant planets to get information on the equation of state ({\\small EOS}), opacities or thermonuclear reactions \\citep[see][for a review]{2006RvMP...78..755R}. Modern ground-based and spatial telescopes equipped with high quality instrumentation are in use or under development ({\\small VLT-VLTI, JWST,} etc.). They provide very accurate data which, after treatment, give access to stellar global parameters (luminosity, radius, mass, effective temperature $T_{\\rm eff}$, gravity $\\log g$, abundances, etc.). On the other hand, seismic data (such as oscillation frequencies or amplitudes) are being obtained in velocity from the ground \\citep[see, e.g.][]{2007CoAst.150..106B} and in photometry by the space missions {\\small MOST} \\citep{2003PASP..115.1023W} and {\\small CoRoT} \\citep{2006corm.book...39M}. In the next decade, valuable observational data are expected. For instance {\\small GAIA} \\citep{ESA2000,2001A&A...369..339P}, to be launched in 2011, will make astrometric measurements, at the micro-arc second level together with photometric and spectroscopic observations of a huge number of stars covering the whole range of stellar masses, compositions and evolution stages while the Kepler mission, to be launched in 2009, will provide the opportunity to make asteroseismic observations on a wide range of stars \\citep{2007CoAst.150..350C}. In the following, I discuss the different aspects of stellar modeling, the problems encountered and the perspectives. ", "conclusions": "" }, "0801/0801.1217_arXiv.txt": { "abstract": "We selected a sample of a dozen blazars which are the prime candidates for simultaneous multi-wavelength observing campaigns in their outburst phase. We searched for optical outbursts, intra-day variability and short term variability in these blazars. We carried out optical photometric monitoring of nine of these blazars in 13 observing nights during our observing run October 27, 2006 $-$ March 20, 2007 by using the 1.02 meter optical telescope equipped with CCD detector and BVRI Johnson broad band filters at Yunnan astronomical observatory, Kunming, China. From our observations, our data favor the hypothesis that three blazars: AO 0235$+$164, S5 0716$+$714 and 3C 279 were in the outburst state; one blazar: 3C 454.3 was in the post outburst state; three blazars: S2 0109$+$224, PKS 0735$+$178 and OJ 287 were in the pre/post outburst state; one blazar: ON 231 was in the low-state; and the state of one blazar: 1ES 2344$+$514 was not known because there is not much optical data available for the blazar to compare with our observations. We observed densely sampled 1534 image frames of these nine blazars. Out of three nights of observations of AO 0235$+$164, intra-day variability was detected in two nights. Out of five nights of observations of S5 0716$+$714, intra-day variability was detected in two nights. In one night of observations of PKS 0735$+$178, intra-day variability was detected. Out of six nights of observations of 3C 454.3, intra-day variability was detected in three nights. No intra-day variability was detected in S2 0109$+$224, OJ 287, ON 231, 3C 279 and 1ES 2344$+$514 in their 1, 4, 1, 2 and 1 nights of observations respectively. AO 0235$+$164, S5 0716$+$714, OJ 287, 3C 279 and 3C 454.3 were observed in more than one night and short term variations in all these blazars were also noticed. From our observations and the available data, we found that the predicted optical outburst with the time interval of $\\sim$ 8 years in AO 0235$+$164 and $\\sim$ 3 years in S5 0716$+$714 have possibly occurred. ", "introduction": "Blazars represent a small subset of the most enigmatic class of radio-loud active galactic nuclei (AGN), exhibiting strong variability at all wavelengths of the whole electromagnetic (EM) spectrum, strong polarization from radio to optical wavelengths, and usually core dominated radio structures. The radiation of blazars at all wavelengths is predominantly nonthermal. This class includes BL Lacertae (BL Lac) objects and flat-spectrum radio quasars (FSRQs). BL Lacs show largely featureless optical continuum. In a unified model of the radio-loud AGN based on the angle between the line of sight and the emitted jet from the source, blazars jet make angle of $\\sim$ 10$^{\\circ}$ from the line of sight (Urry \\& Padovani 1995). The radiation emitted by the plasma, with bulk relativistic motion in the jet oriented at small viewing angles, is affected by relativistic beaming, which in turn implies a shortening of time scales by a factor $\\delta^{-1}$, where $\\delta$ is the Doppler factor. From observations of blazars, it is known that they vary on the diverse time scales. Variability time scales of blazars can be broadly divided into 3 classes viz. intra-day variability (IDV) or micro-variability, short term outbursts and long term trends. Significant variations in flux of a few tenths of magnitude over the course of a day or less is often known as IDV (Wagner \\& Witzel 1995). Short term outburst and long term trends can have time scales range from few weeks to several months and several months to years, respectively. In last about two decades, variability of blazars in radio to optical bands on diverse time scales have been reported in a large number of papers (e.g. Miller et al. 1989; Courvoisier, et al. 1995; Heidt \\& Wagner 1996; Takalo et al. 1996; Sillanp$\\ddot{a}\\ddot{a}$ et al. 1996a, 1996b; Bai, et al. 1998, 1999; Fan et al. 1998, 2002, 2007; Xie, et al. 2002a; Gupta et al. 2004; Ciprini et al. 2003, 2007 and references therein). We selected a sample of a dozen blazars which are prominent candidates for simultaneous multi-wavelength observing campaigns in their outburst phase. The motivation of the present work was to observe these blazars in search for IDV, short term variability and also find out if there is any one being in the outburst state. Blazar emission mechanism in the outburst state and detected IDV is strongly supported by the jet based models of radio-loud AGN. In general, blazar emission in the outburst state is nonthermal Doppler boosted emission from jets (Blandford \\& Rees 1978; Marscher \\& Gear 1985; Marscher et al. 1992, Hughes et al. 1992). There are other models of AGN that can explain the IDV in any type of AGN are optical flares, disturbances or hot spots on the accretion disk surrounding the black hole of the AGN (Mangalam \\& Wiita 1993 and references therein). Models based on the instabilities on accretion disc are mainly supported blazars IDV when the blazar is in the low-state. When a blazar is in the low-state, any contribution from the jets if at all present, is very weak. Recently, it is noticed that, in the low luminosity AGN, accretion disk is radiatively inefficient (Chiaberge et al. 2006; Capetti et al. 2007). So, there will be an alternative way to explain the IDV in the low-state of blazars, in which a weak jet emission will be responsible for the IDV. With this motivation we recently carried out optical photometric observations of the nine blazars: S2 0109$+$224, AO 0235$+$164, S5 0716$+$714, PKS 0735$+$178, OJ 287, ON 231, 3C 279 and 1ES 2344$+$514 in R passband and 3C 454.3 in V and R passbands. The paper is arranged as follows: section 2 describes observations and data analysis method, in section 3 we mentioned our results, discussion and conclusion of the present work is reported in section 4. ", "conclusions": "" }, "0801/0801.1398_arXiv.txt": { "abstract": "We study a 5-dimensional $f({\\cal R})$ brane gravity within the framework of scalar-tensor type theories. We show that such a model predicts, for a certain choice of $f({\\cal R})$ and a spatially flat universe, an exponential potential, leading to an accelerated expanding universe driven solely by the curvature of the bulk space. This result is consistent with the observational data in the cosmological scale. ", "introduction": "The idea that our world might be a brane embedded in a higher dimensional space-time (the bulk) \\cite{1} has been in the mainstream of cosmological investigations in the past few years \\cite{3,4}. This approach differs from the usual Kaluza-Klein idea in that the size of the extra dimensions can be large. The concept of large extra dimensions is discussed phenomenologically in \\cite{5}. An important ingredient of the brane world scenario is that the matter is confined to the brane and the only communication between the brane and bulk is through gravitational interaction or some other dilatonic matter. In general, the matter on the brane leads to a cosmological evolution which is different from the usual evolution governed by the Friedmann equation, that is, in brane cosmology the Hubble parameter on the brane is proportional to the square of energy density \\cite{3,4}. This proportionality is a result of the application of the Israel matching condition which is basically a relation between the extrinsic curvature and the energy-momentum tensor representing matter fields on the brane. Although in brane theories matter fields live on the brane, the possibility of the presence of matter in the form of a scalar field in the bulk has also been investigated in several works. One of the first motivations to introduce a bulk scalar field was to stabilize \\cite{7} the distance between the two branes in the context of the first model introduced by Randall and Sundrum \\cite{1}. A second motivation was the possibility of the resolution of the famous cosmological constant problem \\cite{8}. Several works have studied, in particular, the impact of the presence of a scalar field in the bulk on the cosmological evolution on the brane, without trying to solve the full system of equations in the bulk \\cite{9,10}. In \\cite{12}, the authors have addressed some of the solutions for these equations and studied the corresponding brane evolution. The purpose of the present study is to employ modified gravity \\cite{parry} in the Einstein frame to explain the origin of such a self interacting scalar potential. An interesting observation made a few years ago was that the expansion of our universe is currently undergoing a period of acceleration which is directly measured from the light curves of several hundred type Ia supernovae \\cite{13} and independently from observations of the cosmic microwave background (CMB) by the WMAP satellite \\cite{15} and other CMB experiments \\cite{16}. However, the mechanism responsible for this acceleration is not well understood and many authors introduce a mysterious cosmic fluid, the so called dark energy, to explain this effect \\cite{17}. Recently, it has been shown that such an accelerated expansion could be the result of a modification to the Einstein-Hilbert action \\cite{18} in the framework of DGP brane cosmology. In the present work we study the general form of the Einstein-Hilbert action for any function of the Ricci scalar, $ f({\\cal R})$, in 5 dimensions. This is done in the framework of a scalar-tensor type theory \\cite{20} where a scalar field is minimally coupled to gravity with a self-interacting potential. In this formulation we obtain explicit solutions using conformal transformations, a technique employed in the case of an empty bulk with a cosmological constant \\cite{21} or a bulk with a scalar field, similar to the present work, but with an exponential potential. We present explicit solutions for a particular choice of $ f({\\cal R})$ which predict a similar exponential potential. The organization of the manuscript is as follows: in section 2 we briefly review the scalar-tensor formulation in 5-dimensions and write the full system of equations. In section 3 we consider the cosmological equations for $f({\\cal R})$ gravity which, in the Einstein frame, correspond to a self interacting scalar field with a certain potential. Finally, we study the cosmological evolution on the brane for ${\\cal R}^{m}$ gravity which predicts a power law acceleration in section 4. Conclusions are drawn in the last section. ", "conclusions": "In this manuscript we have obtained explicit solutions in a brane world scenario where an arbitrary function of the Ricci scalar is taken as the bulk Lagrangian. Using a conformal transformation, the action is converted to that of a scalar-tensor type theory with a scalar field. We have shown that with a suitable choice for the function $f({\\cal R})$ and brane tension $\\lambda$, an accelerated expanding universe emerges. The source of this acceleration is not related to an exotic matter but to a scalar field whose origin can be traced back to geometry of the brane and, specifically, to the curvature scalar $\\cal R$ and depends on two free parameters, namely $\\alpha$ and $\\lambda_c$. Hence, an accelerating universe driven by curvature would certainly seem to be a possibility." }, "0801/0801.1167_arXiv.txt": { "abstract": "With the goal to study the physical and chemical evolution of ices in solar-mass systems, a spectral survey is conducted of a sample of 41 low luminosity YSOs ($L\\sim 0.1-10~{\\rm L}_{\\odot}$) using 5--38 \\mum\\ Spitzer Space Telescope and 3--4 \\mum\\ ground-based spectra. The sample is complemented with previously published Spitzer spectra of background stars and with ISO spectra of well studied massive YSOs ($L\\sim 10^5~{\\rm L}_{\\odot}$). This paper focuses on the origin of the prominent absorption features in the 5-8 \\mum\\ spectral region. The long-known 6.0 and 6.85 \\mum\\ bands are detected toward all sources, with the Class 0-type low mass YSOs showing the deepest bands ever observed. In almost all sources the 6.0 \\mum\\ band is deeper, by up to a factor of 3, than expected from the bending mode of pure solid H$_2$O, based on the optical depths of the 3.0 \\mum\\ stretching and 13 \\mum\\ libration modes. The depth and shape variations of the remaining 5--7 \\mum\\ absorption indicate that it consists of 5 independent components, which, by comparison to laboratory studies, must be from at least 8 different carriers. Together with information from the 3-4 \\mum\\ spectra and the additionally detected weak 7.25, 7.40, 9.0, and 9.7 \\mum\\ features it is argued that overlapping bands of simple species are responsible for much of the absorption in the 5-7 \\mum\\ region, at abundances of 1-30\\% for CH$_3$OH, 3-8\\% for NH$_3$, 1-5\\% for HCOOH, $\\sim$6\\% for H$_2$CO, and $\\sim$0.3\\% for HCOO$^-$ with respect to solid H$_2$O. The 6.85 \\mum\\ band likely consists of one or two carriers, of which one is less volatile than H$_2$O because its abundance relative to H$_2$O is enhanced at lower H$_2$O/$\\tau_{9.7}$ ratios. It does not survive in the diffuse interstellar medium (ISM), however. The similarity of the 6.85 \\mum\\ bands for YSOs and background stars indicates that its carrier(s) must be formed early in the molecular cloud evolution. If an NH$_4^+$ salt is the carrier its abundance with respect to solid H$_2$O is typically 7\\%, and low temperature acid-base chemistry or cosmic ray induced reactions must have been involved in its formation. Possible origins are discussed for the carrier of an enigmatic, very broad absorption between 5 and 8 \\mum. It shows large depth variations toward both low- and high-mass YSOs. Weak evidence is found that it correlates with temperature tracers. Finally, all the phenomena observed for ices toward massive YSOs are also observed toward low mass YSOs, indicating that processing of the ices by internal ultraviolet radiation fields is a minor factor in the early chemical evolution of the ices. ", "introduction": "~\\label{sec:intro} The infrared spectra of protostars and obscured background stars show prominent absorption features at 3.0, 4.25, 4.7, 6.0, 6.85, 9.7, and 15 \\mum\\ along with a suite of weaker features (e.g. \\citealt{dhe96}, \\citealt{whi96}, \\citealt{ger99}, \\citealt{sch99}, \\citealt{gib00}, \\citealt{kea01b}, \\citealt{pon03a}, \\citealt{gib04}, \\citealt{kne05}, \\citealt{whi07}; see \\citealt{boo04} for a complete list of features). These are attributed to absorption in the vibrational modes of molecules in ices, except for the 9.7 \\mum\\ band which is mostly due to silicates. At the low temperatures ($T\\leq$20 K) of dense clouds and circum-protostellar environments atoms and molecules freeze out rapidly on dust grains. Grain surface chemistry efficiently forms new, simple species, such as H$_2$O and H$_2$CO (e.g. \\citealt{tie82}). Complex species (e.g. polyoxymethylene [`POM', -(CH$_2$-O)$_{\\rm n}$-], and hexamethylenetetramine ['HMT', C$_6$H$_{12}$N$_4$]) can be formed through the impact of energetic photons or cosmic rays on the ices, as many laboratory studies have shown (e.g. \\citealt{sch93}, \\citealt{ber95}, \\citealt{gre95}, \\citealt{ger96}, \\citealt{moo98}, \\citealt{pal00}, \\citealt{mun04}). Thus far, only the simple species H$_2$O, CO, CO$_2$, CH$_4$, CH$_3$OH, NH$_3$ and the $^{13}$CO and $^{13}{\\rm CO}_2$ isotopes have been positively identified in the ices toward both low and high mass protostars as well as extincted background stars whose lines of sight trace quiescent dense cloud material. Reasonable evidence exists for the presence of HCOOH, OCS and the ions NH$_4^+$, OCN$^-$, and HCOO$^-$ as well, although their existence is sometimes debated because of inaccurate fits with laboratory spectra or the lack of multiple bands for independent confirmation. The identification of the ions in the ices has been particularly controversial. The first ion claimed was OCN$^-$ \\citep{gri87} and was thought to be produced by heavy energetic processing of CO:NH$_3$ ices. Later it was realized that acid-base chemistry in a HNCO:NH$_3$ ice could yield the same products \\citep{nov01}. Acid-base reactions are very efficient and occur at temperatures as low as 10 K \\citep{rau04b}. The created ions are less volatile than neutral species and would be able to form an interstellar salt after the other species have sublimated. The presence of complex species formed by ultraviolet (UV) photon and cosmic ray processing of the simple ices is at least as controversial. Many claims were made \\citep{gre95,gib02}, but the observational evidence is not firm. It is crucial to determine the complexity of the ices in the circumstellar environment of low mass Young Stellar Objects (YSOs), as they may be delivered directly to comets and they may ultimately serve as the source of volatiles in planets. At sublimation fronts in the warm inner regions of disks they are the starting point of a complex gas phase chemistry. The relative abundances of N-, C-, and O-bearing species coming from the ices determine directly the type of chemistry in hot cores (or `hot corinos' for low mass objects; e.g. \\citealt{caz03}). As a step toward better characterizing the molecular content of icy grain mantles, this work presents spectra of a large sample of low mass protostars and background stars obtained with the InfraRed Spectrometer (IRS; \\citealt{hou04}) at the Spitzer Space Telescope \\citep{wer04} in the 5-35 \\mum\\ wavelength range. These data are complemented with spectra at 2-4 \\mum\\ obtained with ground-based facilities. Such full coverage over the near and mid-infrared wavelength ranges is essential in determining the composition of the ices; most species have several vibrational modes and more secure identifications can be made when the features are studied simultaneously. The identification process is aided by studying large samples of sight-lines facilitating correlation of band strengths and shapes with each other and, if known, with physical characteristics along the sight-lines, such as luminosity and evolutionary stage of any heating source. The great sensitivity of the IRS allows for observations of objects with luminosities that are down by orders of magnitude with respect to what could be observed before in the mid-infrared (typically 1 L$_{\\odot}$ versus $10^4$ L$_{\\odot}$). Most data presented in this paper were obtained from spectral surveys of nearby molecular clouds and isolated dense cores in the context of the Spitzer Legacy Program `From Molecular Cores to Planet-Forming Disks' (`c2d'; \\citealt{eva03}). Initial results from this program emphasize the importance of thermal processing in the evolution of the ices, as the ices surrounding the low mass YSO HH~46 IRS are more processed compared to B5 IRS1 \\citep{boo04_2}. Ices toward the edge-on disk CRBR 2422.8-3423 were presented in \\citet{pon05} and also show signs of heating. Finally, mid-infrared spectra of ices toward background stars tracing quiescent cloud material were published in \\citet{kne05}, showing for the first time that the (unknown) carrier of the 6.85 \\mum\\ absorption band is abundant even at the coldest conditions away from star formation. This paper specifically addresses questions on the identification of the 6.0 and 6.85 \\mum\\ bands. These prominent absorption features are still not fully identified, despite being detected nearly 30 years ago with the Kuiper Airborne Observatory \\citep{pue79} and commonly observed toward massive YSOs with the Infrared Space Observatory (ISO; \\citealt{sch96}, \\citealt{kea01b}). The peak position of the 6.85 \\mum\\ band varies dramatically, and is thought to be a function of the ice temperature \\citep{kea01b}. Either one carrier with a pronounced temperature dependence of its absorption bands, or two independent carriers, one much more volatile than the other, must be responsible for the 6.85 \\mum\\ feature. If the former is the case, the ammonium ion (NH$_4^+$) is considered a promising candidate \\citep{sch03}. The 6.0 \\mum\\ band was initially thought to be mainly due to the bending mode of H$_2$O (\\citealt{tie84}; however, see \\citealt{cox89}), but analysis of ISO spectra indicated that this is not the case toward several massive YSOs \\citep{sch96, kea01b}. The depth of the 6.0 \\mum\\ band is in some cases significantly (factor of 2) deeper than that expected from the 3.0 \\mum\\ H$_2$O stretching mode. Part of this `excess' might be due to a strong vibration of the HCOOH molecule. Recent observations of background stars show a small excess ($\\leq$25\\%), of which half could be due to the enhanced band strength of the H$_2$O bending mode in CO$_2$-rich ices \\citep{kne05}. Finally, it was claimed that much of the 6.0 \\mum\\ excess and part of the 6.85 \\mum\\ band is due to highly processed ices, i.e. a mixture of complex species with C--H and O--H bonds produced after irradiation of simple ices \\citep{gib02}. In this case, one would expect the 6.0 \\mum\\ excess to be enhanced in high radiation environments, such as near more evolved protostars. Thus, fundamental questions remain on the identification of the 6.0 and 6.85 \\mum\\ bands, and with the large sample presented in this work the possible answers can be further constrained. Subsequent papers will specifically address the CO$_2$ \\citep{pon08}, CH$_4$ \\citep{obe08}, and NH$_3$ (S. Bottinelli et al., in preparation) species. Further papers in this series will investigate the ices toward background stars behind large clouds (C. Knez et al., in preparation) and isolated cores (A. C. A. Boogert et al., in preparation). In \\S\\ref{sec:sou} the source sample is presented, while the observations and data reduction of the ground-based L-band and Spitzer 5-38 \\mum\\ spectra are described in \\S\\ref{sec:obs}. In \\S\\ref{sec:res} the continuum level for the multitude of absorption features is determined, and subsequently the best value for the H$_2$O column density derived, followed by an empirical decomposition of the 5-7 \\mum\\ absorption complex. The positively identified species, or correlations of the components with other observables are discussed in \\S\\ref{sec:id}. Finally, the abundances of the simple species, and constraints on the carriers of unidentified components are discussed in \\S\\ref{sec:dis}. \\begin{deluxetable*}{lllllrlll} \\tabletypesize{\\scriptsize} \\tablecolumns{9} \\tablewidth{0pc} \\tablecaption{Source Sample~\\label{t:sample}} \\tablehead{ \\colhead{Source}& \\colhead{RA\\tablenotemark{a}} & \\colhead{Dec\\tablenotemark{a}} & \\colhead{Cloud} & \\colhead{Type\\tablenotemark{b}} & \\colhead{$\\alpha_{\\rm IR}$\\tablenotemark{c}} & \\colhead{Obs. ID\\tablenotemark{d}} & \\colhead{Module\\tablenotemark{l}} & \\colhead{L-band\\tablenotemark{m}}\\\\ \\colhead{ }& \\colhead{J2000} & \\colhead{J2000} & \\colhead{ } & \\colhead{ } & \\colhead{ } & \\colhead{ } & \\colhead{ } & \\colhead{ }\\\\} \\startdata L1448 IRS1 & 03:25:09.44 & +30:46:21.7 & Perseus & low & 0.34 & \\dataset{0005656832} & SL,SH,LH & NIRSPEC \\\\ IRAS 03235+3004 & 03:26:37.45 & +30:15:27.9 & Perseus & low & 1.44 & \\dataset{0009835520} & SL,LL & NIRSPEC \\\\ IRAS 03245+3002 & 03:27:39.03 & +30:12:59.3 & Perseus & low & 2.70 & \\dataset{0006368000} & SL,SH & \\\\ L1455 SMM1 & 03:27:43.25 & +30:12:28.8 & Perseus & low & 2.41 & \\dataset{0015917056} & SL,SH,LL1 & \\\\ RNO 15 & 03:27:47.68 & +30:12:04.3 & Perseus & low &$-$0.21 & \\dataset{0005633280} & LL1,SL,SH,LH& NIRSPEC \\\\ L1455 IRS3 & 03:28:00.41 & +30:08:01.2 & Perseus & low & 0.98 & \\dataset{0015917568} & SL,SH,LL1 & \\\\ IRAS 03254+3050 & 03:28:34.51 & +31:00:51.2 & Perseus & low & 0.90 & \\dataset{0011827200} & LL1,SL,SH,LH& NIRSPEC \\\\ IRAS 03271+3013 & 03:30:15.16 & +30:23:48.8 & Perseus & low & 2.06 & \\dataset{0005634304} & LL1,SL,SH,LH& NIRSPEC \\\\ IRAS 03301+3111 & 03:33:12.85 & +31:21:24.2 & Perseus & low & 0.51 & \\dataset{0005634560} & SL,SH,LH & NIRSPEC \\\\ B1-a & 03:33:16.67 & +31:07:55.1 & Perseus & low & 1.87 & \\dataset{0015918080} & SL,SH,LL1 & NIRSPEC \\\\ B1-c & 03:33:17.89 & +31:09:31.0 & Perseus & low & 2.66 & \\dataset{0013460480} & SL,SH,LL1 & \\\\ B1-b & 03:33:20.34 & +31:07:21.4 & Perseus & low & 0.68 & \\dataset{0015916544} & SL,LL & \\\\ B5 IRS3 & 03:47:05.45 & +32:43:08.5 & Perseus & low & 0.51\\tablenotemark{k}& \\dataset{0005635072} & LL1,SL,SH,LH& NIRSPEC \\\\ B5 IRS1\\tablenotemark{e} & 03:47:41.61 & +32:51:43.8 & Perseus & low & 0.78\\tablenotemark{k}& \\dataset{0005635328} & SL,SH,LH & NIRSPEC \\\\ L1489 IRS\\tablenotemark{f} & 04:04:43.07 & +26:18:56.4 & Taurus & low & 1.10 & \\dataset{0003528960} & SL,SH,LH & NIRSPEC \\\\ IRAS 04108+2803B\\tablenotemark{f} & 04:13:54.72 & +28:11:32.9 & Taurus & low & 0.90 & \\dataset{0003529472} & SL,SH,LH & NIRSPEC \\\\ HH~300\\tablenotemark{f} & 04:26:56.30 & +24:43:35.3 & Taurus & low & 0.79 & \\dataset{0003530752} & SL,SH,LH & NIRSPEC \\\\ DG Tau B\\tablenotemark{f} & 04:27:02.66 & +26:05:30.5 & Taurus & low & 1.16 & \\dataset{0003540992} & SL,SH,LH & NIRSPEC \\\\ HH~46 IRS\\tablenotemark{e} & 08:25:43.78 &$-$51:00:35.6 & HH~46 & low & 1.70 & \\dataset{0005638912} & SL,SH,LH & ISAAC \\\\ IRAS 12553-7651 & 12:59:06.63 &$-$77:07:40.0 & Cha & low & 0.76 & \\dataset{0009830912} & LL1,SL,SH,LH& \\\\ IRAS 13546-3941 & 13:57:38.94 &$-$39:56:00.2 & BHR92 & low & $-$0.06 & \\dataset{0005642752} & SL,SH,LH & \\\\ IRAS 15398-3359 & 15:43:02.26 &$-$34:09:06.7 & B228 & low & 1.22 & \\dataset{0005828864} & SL,SH,LL1 & \\\\ Elias 29\\tablenotemark{g} & 16:27:09.42 &$-$24:37:21.1 & Oph & low & 0.53\\tablenotemark{k}& 26700814 & SWS01 sp3 & \\\\ CRBR 2422.8-3423\\tablenotemark{h} & 16:27:24.61 &$-$24:41:03.3 & Oph & low & 1.36 & \\dataset{0009346048} & SL,SH,LH & NIRSPEC \\\\ RNO 91 & 16:34:29.32 &$-$15:47:01.4 & L43 & low & 0.03 & \\dataset{0005650432} & SL,SH,LH & ISAAC \\\\ IRAS 17081-2721 & 17:11:17.28 &$-$27:25:08.2 & B59 & low & 0.55 & \\dataset{0014893824} & SL,SH,LL1 & NIRSPEC \\\\ SSTc2dJ171122.2-272602 & 17:11:22.16 &$-$27:26:02.3 & B59 & low & 2.26 & \\dataset{0014894336} & SL,LL & \\\\ 2MASSJ17112317-2724315 & 17:11:23.13 &$-$27:24:32.6 & B59 & low & 2.48 & \\dataset{0014894592} & SL,LL & NIRSPEC \\\\ EC 74 & 18:29:55.72 & +01:14:31.6 & Serpens & low & $-$0.25 & \\dataset{0009407232} & SL,SH,LH & NIRSPEC \\\\ EC 82 & 18:29:56.89 & +01:14:46.5 & Serpens & low & 0.38\\tablenotemark{k}& \\dataset{0009407232} & SL,SH,LH & \\\\ SVS 4-5 & 18:29:57.59 & +01:13:00.6 & Serpens & low & 1.26 & \\dataset{0009407232} & SL,SH,LH & ISAAC \\\\ EC 90 & 18:29:57.75 & +01:14:05.9 & Serpens & low & $-$0.09 & \\dataset{0009828352} & SL,SH,LH & \\\\ EC 92 & 18:29:57.88 & +01:12:51.6 & Serpens & low & 0.91 & \\dataset{0009407232} & SL,SH,LH & NIRSPEC \\\\ CK4 & 18:29:58.21 & +01:15:21.7 & Serpens & low &$-$0.25 & \\dataset{0009407232} & SL,SH,LH & \\\\ R CrA IRS 5 & 19:01:48.03 &$-$36:57:21.6 & CrA & low & 0.98 & \\dataset{0009835264} & SL,SH,LL1 & ISAAC \\\\ HH 100 IRS\\tablenotemark{i} & 19:01:50.56 &$-$36:58:08.9 & CrA & low & 0.80 & 52301106 & SWS01 sp4 & \\\\ CrA IRS7 A & 19:01:55.32 &$-$36:57:22.0 & CrA & low & 2.23 & \\dataset{0009835008} & SL,SH,LH & ISAAC \\\\ CrA IRS7 B & 19:01:56.41 &$-$36:57:28.0 & CrA & low & 1.63 & \\dataset{0009835008} & SL,SH,LH & ISAAC \\\\ CrA IRAS32 & 19:02:58.69 &$-$37:07:34.5 & CrA & low & 2.15 & \\dataset{0009832192} & SL,SH,LL1 & \\\\ L1014 IRS & 21:24:07.51 & +49:59:09.0 & L1014 & low & 1.28 & \\dataset{0012116736} & SL,LL & NIRSPEC \\\\ IRAS 23238+7401 & 23:25:46.65 & +74:17:37.2 & CB 244 & low & 0.95 & \\dataset{0009833728} & SL,SH,LH & NIRSPEC \\\\ & & & & & & & & \\\\ W3 IRS5\\tablenotemark{i} & 02:25:40.54 &$+$62:05:51.4 & & high & 3.53 & 42701302 & SWS01 sp3 & \\\\ MonR2 IRS3\\tablenotemark{i} & 06:07:47.8 &$-$06:22:55.0 & & high & 1.66 & 71101712 & SWS01 sp3 & \\\\ GL989\\tablenotemark{i} & 06:41:10.06 &$+$09:29:35.8 & & high & 0.52 & 71602619 & SWS01 sp3 & \\\\ W33A\\tablenotemark{i} & 18:14:39.44 &$-$17:52:01.3 & & high & 1.92 & 32900920 & SWS01 sp4 & \\\\ GL7009S\\tablenotemark{i} & 18:34:20.91 &$-$05:59:42.2 & & high & 2.52 & 15201140 & SWS01 sp3 & \\\\ GL2136\\tablenotemark{i} & 18:22:26.32 &$-$13:30:08.2 & & high & 1.48 & 33000222 & SWS01 sp3 & \\\\ S140 IRS1\\tablenotemark{i} & 22:19:18.17 &$+$63:18:47.6 & & high & 1.57 & 22002135 & SWS01 sp4 & \\\\ NGC7538 IRS9\\tablenotemark{i} & 23:14:01.63 &$+$61:27:20.2 & & high & 2.31 & 09801532 & SWS01 sp2 & \\\\ & & & & & & & & \\\\ Elias 16\\tablenotemark{k} & 04:39:38.88 \t &$+$26:11:26.6 & Taurus & bg & - & \\dataset{0005637632} & SL,SH & \\\\ EC 118\\tablenotemark{k} & 18:30:00.62 \t &$+$01:15:20.1 & Serpens & bg & - & \\dataset{0011828224} & SL,SH & \\\\ \\enddata \\tablenotetext{a}{Position used in Spitzer/IRS observations} \\tablenotetext{b}{Source type: 'low'=low mass YSO, 'high'=massive YSO, 'bg'=background star} \\tablenotetext{c}{Broad-band spectral index as defined in Eq.~\\ref{eq:alpha}} \\tablenotetext{d}{AOR key for Spitzer and TDT number for ISO observations} \\tablenotetext{e}{Published previously in Boogert et al. 2004} \\tablenotetext{f}{Published previously in Watson et al. 2004} \\tablenotetext{g}{Published previously in Boogert et al. 2000} \\tablenotetext{h}{Published previously in Pontoppidan et al. 2005} \\tablenotetext{i}{Published previously in Keane et al. 2001} \\tablenotetext{j}{Published previously in Knez et al. 2005} \\tablenotetext{k}{$\\alpha$ enhanced due to foreground extinction. Exclusion $K_{\\rm s}$-band flux gives much lower $\\alpha$: $-$0.16 (Elias 29), $-0.02$ (B5 IRS1), 0.18 (B5 IRS3), and 0.38 (EC 82)} \\tablenotetext{l}{Spitzer/IRS modules used: SL=Short-Low (5-14 \\mum, $R\\sim100$), LL=Long-Low (14-34 \\mum, $R\\sim100$), SH=Short-High (10-20 \\mum, $R\\sim600$), LH=Long-High (20-34 \\mum, $R\\sim600$); ISO SWS modes used (2.3-40 \\mum): SWS01 speed 1 ($R\\sim250$), speed 2 ($R\\sim250$), speed 3 ($R\\sim400$), speed 4 ($R\\sim800$)} \\tablenotetext{m}{Complementary ground-based L-band observations with Keck/NIRSPEC or VLT/ISAAC} \\end{deluxetable*} ", "conclusions": "~\\label{sec:concl} The present work extends the study of ices over the full 3-20 \\mum\\ wavelength range from the previously well studied massive YSOs ($\\sim10^5$L$_{\\odot}$) to low mass YSOs ($\\sim$1L$_{\\odot}$). The following new insights were obtained: \\begin{itemize} \\item Absorption features are detected at 6.0 and 6.85 \\mum\\ in all sources, including high- and low-mass YSOs and background stars tracing quiescent cloud material. An empirical decomposition shows that the 5-7 \\mum\\ absorption complex consists, in addition to the H$_2$O bending mode, of a combination of at least 5 independent absorption components. \\item In a subset of sources additional weak features are detected at 7.25, 7.40, and at 9.0 and 9.7 \\mum\\ on top of the prominent Si-O stretching band of silicates. These features are associated with CH$_3$OH, HCOOH, NH$_3$, and possibly HCOO$^-$ and indicate abundances of 1-30\\%, 1-5\\%, 3-8\\%, and 0.3\\% with respect to H$_2$O, respectively. The large source-to-source solid CH$_3$OH abundance variations are likely a result of the conditions at the time of grain surface chemistry. \\item Component C1 (5.7-6.0 \\mum) is mostly explained by solid HCOOH and H$_2$CO at abundances with respect to solid H$_2$O of typically 1-5\\% and $\\sim$6\\%, respectively. \\item Component C2 (6.0-6.4 \\mum) also likely arises from a blend of several species. Solid NH$_3$ can account for 10-50\\% of the absorption. A similar amount can be attributed to absorption by monomers, dimers, and small multimers of H$_2$O mixed with a substantial amount of CO$_2$. The long-wavelength side of C2 is potentially due to anions produced by acid-base chemistry (e.g. HCOO$^-$) or energetic processing. Indeed, a weak correlation with C4 suggests the carriers of these bands (possibly salts) are related. \\item Components C3 and C4, the 6.85 \\mum\\ band, show the same characteristics for low mass YSOs and background stars as was previously found for massive YSOs by \\citet{kea01b}. Their ratio is empirically found to be dependent on dust temperature. The carrier of C4 is likely less volatile than that of C3 and than H$_2$O, but it is not observed in the diffuse ISM. These characteristics are also consistent with both components being due to NH$_4^+$ (ammonium ion). The detection of strong 6.85 \\mum\\ bands toward deeply embedded YSOs and background stars requires a production at low temperature (acid-base chemistry or processing by cosmic rays), given the lack of heating sources and stellar UV fields. \\item The origin of the very broad component C5 (5-8 \\mum) is least understood. It is quite strong in several low and high mass YSOs, absent in others and so far undetected toward background stars. It is possibly related to thermal processing, as a weak correlation with the ratio of the polar/apolar CO components is observed and its shape resembles that of warm H$_2$O. A blend of absorption by ions, as proposed by \\citep{sch03}, which would require both heating and energetic processing of the ices, cannot be excluded. The latter could also lead to an organic residue, whose shape resembles that of the C5 component as well. \\item Weak correlations are found between the absorption components in the 5-8 \\mum\\ range and tracers of physical conditions. A more thorough understanding of the conditions, the source geometry and local radiation fields in specific lines of sight is required to further constrain the nature of the carriers (in particular for components C2-C5) and the importance of thermal and energetic processing. To what degree does scattering play a role? What are the dust temperatures, cosmic ray fluxes, UV radiation fields and time scales as a function of distance along the line of sight? Measurements of source fluxes above wavelengths of 100 \\mum\\ with the Herschel Space Observatory will be valuable, as well as ground-based measurements of the 4.62 \\mum\\ band of OCN$^-$ in more lines of sight. In addition, further laboratory work, in particular on the effect of cosmic rays on the ices, especially for the identification of components C3 and C4, is required. Finally, study of the 5-8 \\mum\\ features toward a larger sample of background stars is crucial to make further progress in the identification and processing history of interstellar ices. \\end{itemize}" }, "0801/0801.1956_arXiv.txt": { "abstract": "Continuous observations were obtained of active region 10953 with the Solar Optical Telescope (SOT) on board the \\emph{Hinode} satellite during 2007 April 28 to May 9. A prominence was located over the polarity inversion line (PIL) in the south-east of the main sunspot. These observations provided us with a time series of vector magnetic fields on the photosphere under the prominence. We found four features: (1) The abutting opposite-polarity regions on the two sides along the PIL first grew laterally in size and then narrowed. (2) These abutting regions contained vertically-weak, but horizontally-strong magnetic fields. (3) The orientations of the horizontal magnetic fields along the PIL on the photosphere gradually changed with time from a normal-polarity configuration to a inverse-polarity one. (4) The horizontal-magnetic field region was blueshifted. These indicate that helical flux rope was emerging from below the photosphere into the corona along the PIL under the pre-existing prominence. We suggest that this supply of a helical magnetic flux into the corona is associated with evolution and maintenance of active-region prominences. ", "introduction": "Solar prominences are cool material (10$^4$ K) floating in the corona (10$^6$ K) and located over polarity inversion lines (PILs) of the photosphere. It is known that prominences are supported by coronal magnetic fields against gravity (see references in Martin 1998). Such magnetic fields have dipped shapes, and the cool plasma is sitting at the bottom of the dips (Kippenhahn \\& Schl\\\"{u}ter 1957). Moreover, many observations of prominences show inverse polarity (Leroy et al. 1984; Tandberg-Hanssen 1995; Lites 2005), which means that the direction of horizontal magnetic fields is toward the positive polarity side from the negative polarity side. Hence, it is often thought that magnetic fields in prominences have helical structures (e.g., Kuperus \\& Tandberg-Hanssen 1967; Kuperus \\& Raadu 1974; Hirayama 1985). This topology of magnetic fields is suggested from observations of erupting prominences, or coronal mass ejections (e.g., Dere et al. 1999; Ciaravella et al. 2000; Low 2001). How is such a helical magnetic field created in association with prominences in the corona ? Two theories have been discussed: flux rope models (e.g., Rust \\& Kumar 1994; Low \\& Hundhausen 1995; Low 1996; Lites 2005; Zhang \\& Low 2005) and sheared-arcade models (e.g., Pneuman 1983; van Ballegooijen \\& Martens 1989, 1990; Antiochos et al. 1994; DeVore \\& Antiochos 2000; Martens \\& Zwaan 2001; Aulanier et al. 2002; Karpen et al. 2003; Mackay \\& van Ballegooijen 2005, 2006). In the former model, an originally-twisted flux rope emerges from below the photosphere into the corona. The helical structure is thought to be made by the convection in the solar interior. In the latter one, potential fields in the corona are sheared by the photospheric motion along PILs. As a result, magnetic reconnection occurs among the sheared fields, and then helical fields are constructed in the corona. Both models have the same final configuration, but the processes leading to it are quite different. However, we have no clear observational evidence to support either model, although emerging twisted magnetic fields related to flares have already been observed (e.g., Tanaka 1991; Lites et al. 1995; Leka et al. 1996; Ishii et al. 1998). We have had difficulties due to seeing and discontinuity of observing time with ground-based observations in revealing the mechanism of prominence formation. The \\emph{Hinode} satellite (Kosugi et al. 2007) has a sun-synchronous polar orbit, so that we can observe the Sun continuously without interruption by a spacecraft night. The Solar Optical Telescope (SOT, Tsuneta et al. 2007; Suematsu et al. 2007; Ichimoto et al. 2007; Shimizu et al. 2007) on board \\emph{Hinode} has the Spectro-Polarimeter (SP), which provides long sequences of vector magnetic field measurements at high spatial resolution. We have an observation of one active region that had a pre-existing prominence to investigate the evolution of its associated magnetic fields on the photosphere. We report on the results and findings from this observation in this letter. ", "conclusions": "The \\emph{Hinode}/SP observations presented here offer clear evidence of an emerging helical flux rope along the PIL under an active-region prominence: The {\\it weak-field region} became first wider and then narrower during the emergence of the flux rope, while the orientation of the horizontal magnetic fields observed at the photosphere changes from the normal-polarity to the inverse-polarity configuration. We note that we distinguish the {\\it weak-field region} from the filament channel since the {\\it weak-field region} appears temporarily in the filament channel. Cheung et al. (2007) observed similar polarity reversal in their radiative MHD simulation. Lites (2005) pointed out in a similar evolutionary event that the abutting opposite polarities may spatially separate under an active-region prominence although the cadence of that observation was one day and the resolution was much lower than that of ours. Gaizauskas et al. (1997) also reported that elongated voids appeared along a filament channel seen in off-band H$\\alpha$ a few days before prominence formation. Our results with \\emph{Hinode}/SOT suggest that such broadening is associated with emergence of the horizontal flux tubes. We obtain the following physical parameters from the ME inversion of the SP data for our emerging flux rope event. The {\\it weak-field region} had a width of up to 10,000 km, a duration of one and a half days, and an upward velocity of up to 300 m $s^{-1}$. Hence, the diameter of the flux tube is roughly estimated to be 39,000 km if the maximum upward velocity continues for one and a half days. This estimation is consistent with both the maximum width of the {\\it weak-field region} and typical height of active-region prominences on the limb. The filling factor and strength of horizontal magnetic fields are estimated to be 0.15 and 650 Gauss on average in the {\\it weak-field region}, respectively. When this flux is supplied into the corona, the average strength of the magnetic fields is 100 Gauss. This value is consistent with active-region prominences according to previous measurements of magnetic fields (e.g., Tandberg-Hanssen and Malville 1974; Wiehr and Stellmacher 1991; Casini et al. 2003; Okamoto et al. 2007), and this flux rope is strong enough to support prominences. Therefore, we suggest that this emergence of the helical magnetic flux rope is associated with evolution and maintenance of the prominence. This is the only observation made so far of an active-region prominence, and we do not rule out the sheared arcade model (see Introduction). It will be useful to observe quantitatively both active-region and quiescent prominences with the SP to determine whether the event reported in this paper is common in the solar atmosphere. We suggest that the above observational evidence must be closely related to the evolution of a prominence via the emergence of a twisted magnetic flux rope. The H$\\alpha$ prominence was already seen before the start of the emerging episode reported here. The appearance of the pre-existing prominence was considerably changed, but after the emergence of the flux rope, the prominence became stable. We will report the relationship between the emerging helical flux rope and the prominence activity in a subsequence paper. \\ The authors thank B. C. Low for useful comments. \\emph{Hinode} is a Japanese mission developed and launched by ISAS/JAXA, with NAOJ as domestic partner and NASA and STFC (UK) as international partners. It is operated by these agencies in co-operation with ESA and NSC (Norway). This work was carried out at the NAOJ Hinode science center, which was supported by the Grant-in-Aid for Creative Scientific Research ``The Basic Study of Space Weather Prediction'' from MEXT, Japan (Head Investigator: K. Shibata), generous donation from the Sun Microsystems Inc., and NAOJ internal funding." }, "0801/0801.2572_arXiv.txt": { "abstract": "The existence of a primordial magnetic field (PMF) would affect both the temperature and polarization anisotropies of the cosmic microwave background (CMB). It also provides a plausible explanation for the possible disparity between observations and theoretical fits to the CMB power spectrum. Here we report on calculations of not only the numerical CMB power spectrum from the PMF, but also the correlations between the CMB power spectrum from the PMF and the primary curvature perturbations. We then deduce a precise estimate of the PMF effect on all modes of perturbations. We find that the PMF affects not only the CMB TT and TE modes on small angular scales, but also on large angular scales. The introduction of a PMF leads to a better fit to the CMB power spectrum for the higher multipoles, and the fit at lowest multipoles can be used to constrain the correlation of the PMF with the density fluctuations for large negative values of the spectral index. Our prediction for the BB mode for a PMF average field strength $|B_\\lambda| =4.0$ nG is consistent with the upper limit on the BB mode deduced from the latest CMB observations. We find that the BB mode is dominated by the vector mode of the PMF for higher multipoles. We also show that by fitting the complete power spectrum one can break the degeneracy between the PMF amplitude and its power spectral index. ", "introduction": "Magnetic fields in clusters of galaxies have been observed \\cite{Kronberg:1992pp,Wolfe:1992ab, Clarke:2000bz,Xu:2005rb} with a strength of $0.1-1.0~\\mu$ G. The existence of a primordial magnetic field (PMF) of order 1 nG whose field lines collapse as structure forms is one possible explanation for such magnetic fields in galactic clusters. The origin and detection of the PMF is, hence, a subject of considerable interest in modern cosmology. Moreover, the PMF could influence a variety of phenomena in the early universe \\cite{Grasso:2000wj} such as the cosmic microwave background (CMB) \\cite{ Challinor:2005ye, Dolgov:2005ti, Gopal:2005sg, Kahniashvili:2005xe, Kosowsky:2004zh, Lewis:2004ef, Mack:2001gc, Subramanian:1998fn, Subramanian:2002nh, Yamazaki:2005yd, Yamazaki:2004vq, Yamazaki:2006bq, Yamazaki:2006ah}, or the matter density field \\cite{ Giovannini:2004aw, Tashiro:2005hc, Tsagas:1999ft, Yamazaki:2006mi}. Temperature and polarization anisotropies in the CMB provide very precise information on the physical processes in operation during the early universe (WMAP \\cite{Spergel:2006hy,Hinshaw:2006ia,Page:2006hz}, ACBAR \\cite{Kuo:2006ya}, CBI \\cite{Readhead:2004gy,Sievers:2005gj}, DASI \\cite{Leitch:2004gd}, BOOMERANG \\cite{Jones:2005yb}, and VSA \\cite{Dickinson:2004yr}). The CMB power spectrum from ACBAR and CBI, has indicated a potential discrepancy between these observations at higher multipoles $\\ell \\ge 2000$ and the best-fit cosmological model to the WMAP power spectrum. A straightforward extension of the fit \\cite{Spergel:2006hy} to the WMAP data predicts a rapidly declining power spectrum in the large multipole range due to the finite thickness of the photon last scattering surface and the Silk damping effect. The ACBAR and CBI experiments, however, indicate continued power up to $\\ell \\sim 4000$. This discrepancy is difficult to account for by a simple retuning of cosmological parameters or by the Sunyev-Zeldovich effect \\cite{Spergel:2006hy,Komatsu02,Bond05}. Among other possible explanations, an inhomogeneous cosmological magnetic field generated before the CMB last-scattering epoch provides a plausible mechanism \\cite{Bamba:2004cu} to produce excess power at high multipoles. Such a field excites an Alfven-wave mode in the primordial baryon-photon plasma and induces small rotational velocity perturbations. Since this mode can survive on scales below those at which Silk damping occurs during recombination \\cite{Jedamzik:1996wp,Subramanian:1998fn}, it could be a new source of the CMB anisotropies on small angular scales. The present work, therefore, is an attempt to more precisely study the evolution of cosmological perturbations with a PMF. Previous work \\cite{ Durrer:1999bk, Seshadri:2000ky, Campanelli:2004pm, Challinor:2005ye, Dolgov:2005ti, Gopal:2005sg, Kahniashvili:2005xe, Kosowsky:2004zh, Lewis:2004ef, Mack:2001gc, Subramanian:1998fn, Subramanian:2002nh, Yamazaki:2005yd, Yamazaki:2004vq, Yamazaki:2006bq, Yamazaki:2006ah} has shown that one can obtain information about the PMF from the CMB temperature anisotropies and polarization. However, in those works attention was only given to a subset of the modes of the CMB anisotropies. In the present work, therefore, we study the comprehensive effect of the PMF on all modes of the CMB perturbations. Furthermore, in order to clarify the role of the PMF in the CMB, we take into consideration the possible correlation between the CMB fluctuations induced by the PMF and those due to primordial curvature and tensor perturbations. Also, we numerically evaluate the CMB power spectrum from the stochastic PMF and thereby avoid recourse to analytic approximations. In this article, we use adiabatic initial conditions for the evolution of primary density perturbations and consider isocurvature (isothermal) initial conditions when estimating effects on the CMB anisotropy induced by the PMF \\cite{Lewis:2004ef,Giovannini:2006gz}. Throughout this article we fix the best fit cosmological parameters of the $\\Lambda$CDM + Tensor model as follows \\cite{Spergel:2006hy}: $h=0.792$, $\\Omega_bh^2=0.02336$, $\\Omega_m h^2=0.1189$, $n_S=0.987$, $r=0.55$, $n_T=-r/8= -0.069$, and $\\tau_c=0.091$ in flat universe models, where $h$ denotes the Hubble parameter in units of 100 km s$^{-1}$Mpc$^{-1}$, $\\Omega_b$ and $\\Omega_m$ are the baryon and cold dark matter densities in units of the critical density, $n_S$ is the spectral index of the primordial scalar fluctuations, $r$ the ratio of the amplitude of the tensor fluctuations to the scalar potential fluctuations, $n_T$ is the spectral index of the primordial tensor fluctuations, and $\\tau_c$ is the optical depth for Compton scattering. ", "conclusions": "We have explored effects of a PMF on the CMB for the allowed PMF parameters which were deduced in our previous work \\cite{Yamazaki:2006bq} (i.e.~$B_\\lambda < 10~ \\mathrm{nG~and~} n_\\mathrm{B} < -2.4$) . The upper panel of Figure 2 illustrates the CMB temperature and polarization anisotropies from the PMF for scalar, vector and tensor modes for the case when $B_\\lambda = 4.0$ nG and $n_\\mathrm{B} = -2.9$ or $-2.5$ as labeled. The scalar mode dominates for lower $\\ell$ of the TT and TE modes as shown by Giovannini \\cite{Giovannini:2006gz}. In particular, it is comparable in power to the primary TT mode for ($B_\\lambda, n_\\mathrm{B}) = (4.0 ~{\\rm nG}, -2.9$). We note that the curves with ($B_\\lambda, n_\\mathrm{B}, s^\\mathrm{(S)},s^\\mathrm{(T)} ) = (4.0 ~{\\rm nG}, -2.5, 1,1$) give the best fits to the observed power spectra in both the regions of high and low $\\ell$. This result is complementary to and consistent with the nucleosynthesis constraints derived in \\cite{Caprini:2001nb} as shown in \\cite{Yamazaki:2007mm}. For illustration, let us assume the same scale-invariant power spectrum for both the PMF and the primordial curvature perturbations. In this case, the ratio of the density and velocity perturbations induced by the primordial curvature perturbation to those by the PMF is proportional to $k^2$. Therefore, the temperature anisotropies from the PMF are larger for lower $\\ell$ compared with those from primordial curvature perturbations. Furthermore, the power of the CMB temperature anisotropies from the PMF for lower $\\ell$ depends not only on $B_\\lambda$ but also strongly on $n_\\mathrm{B}$ [Panel (1a) of Fig.2]. Note, that the magnetic field which is continually sourcing fluctuations does not spoil the phase coherence (cf.~cosmic defects). The basic behavior of acoustic oscillations is affected by the pressure of the fluid $kc_s\\delta^\\mathrm{(S)}_b$ and the potential $k\\phi$. Since the pressure of the PMF is sufficiently less than the thermal fluid pressure, the PMF dose not affect the phase coherence significantly. The PMF also affects the CMB power spectrum on small angular scales for two reasons. First, the PMF energy density fluctuations depend only on the scale factor $a$ and can survive below Silk damping scale. Therefore, the PMF continues to source the fluctuations through the Lorentz force even below the Silk damping scale. Second, the vector mode from the PMF can be larger than the scalar modes both from the PMF and primordial perturbations at small scales. This is because, after horizon crossing, the latter cannot grow due to the photon pressure leading to acoustic oscillations, while the former can keep growing inside the cosmic horizon. This means that, for higher $\\ell$, the vector mode dominates the temperature anisotropies of the CMB over the scalar and tensor modes from the PMF and the contribution from primary anisotropies. The integrated amplitude of the gravitational waves from the PMF can be negligible after horizon crossing. This is because the homogeneous solution begins to oscillate inside the horizon and decay rapidly \\cite{ 1979ZhPmR..30..719S, 1982PhLB..115..189R, 1985SvA....29..607P, Pritchard:2004qp}. Consequently, gravity waves only affect the anisotropy spectrum on scales larger than the horizon at recombination. The tensor mode from the PMF therefore decreases at higher $\\ell$. Thus, the vector mode of the CMB polarization from the PMF dominates for higher $\\ell$ [Panel (1d) of Fig.~2] \\cite{Seshadri:2000ky,Lewis:2004ef} \\footnote{We do not consider the magnetic-field-related BB polarization coming from the Faraday rotation effect \\cite{Campanelli:2004pm,Kosowsky:2004zh}.} In the primary spectrum for higher $\\ell$ the absolute value of the EE mode from the PMF is relatively-small [Panel (1c) of Fig.~2] compared to the TT mode [Panel (1a) of Fig.~2]. Hence, even though the EE mode of the primary spectrum damps less than the TT mode, the EE mode remains much smaller than the TT mode at higher $\\ell$ in the final power spectrum. For the TE mode [Panel (1b) of Fig.~2], the contribution from the PMF vector and scalar modes can be comparable to that from the primordial curvature perturbations for higher $\\ell$. Except for a small dip region near $\\ell = 40$, the scalar mode is always small compared to the primary spectrum. Regarding the BB mode, we note that the BB mode signal described in this paper is due to magnetic-field-induced CMB fluctuations (with the peak around $\\ell \\sim 2000$ as in \\cite{Seshadri:2000ky,Lewis:2004ef}). We do not include magnetic-field related BB polarization coming from the Faraday rotation effect (with the peak around $\\ell > 15000$) discussed in \\cite{Campanelli:2004pm,Kosowsky:2004zh}. In our model, the BB mode from the PMF can dominate for $\\ell \\gtrsim 200$ if $B_{\\rm \\lambda}\\gtrsim 2.0$ nG. A potential problem in attempting to detect this signal on such angular scales, therefore, is the contamination from gravitational lensing which converts the dominant EE power into the BB mode \\cite{Pritchard:2004qp}. However, since we already know quite accurately what the spectrum of the lensing signal must be, we can subtract its power directly. After removing the foreground effect, the BB mode from the PMF effect dominates for higher $\\ell$ even for PMF parameters allowed by the CMB temperature constraint \\cite{Yamazaki:2004vq,Yamazaki:2006bq,Caprini:2001nb}. Note, that the there is a change of scale between panels 1c and 1b. The BB mode and EE modes are of comparable magnitude. In Panel 2 of Fig. 2 we depict the CMB temperature and polarization anisotropies in the presence of a PMF taking into account the correlations. Since we obtain the temperature and polarization anisotropies of the CMB with the PMF from isocurvature initial conditions, the phase of the CMB perturbation with the PMF is different by $\\pi /2$ from those without the PMF (on the adiabatic initial condition). The first, third, and odd numbered peaks of the scalar mode of the CMB perturbations rise for a positive correlation between the PMF and the energy density perturbations, while they are suppressed for a negative correlation. These are compared in Panel 2 of Figure 2 with the observed power spectra (WMAP \\cite{Hinshaw:2006ia,Page:2006hz}, ACBAR \\cite{Kuo:2006ya}, CBI \\cite{Readhead:2004gy,Sievers:2005gj}, DASI \\cite{Leitch:2004gd}, BOOMERANG \\cite{Jones:2005yb}, and VSA \\cite{Dickinson:2004yr}). Panels (2a) and (2b) of Fig.~2 show clearly that the power spectral index of the PMF, $n_\\mathrm{B}$, is more effectively constrained from CMB observations for lower $\\ell$ than those for higher $\\ell$. The models with higher $n_\\mathrm{B}$ give better fits to the observations than those with lower $n_\\mathrm{B}$ for the lower $\\ell$ regions of the TT and TE modes. Furthermore, there is no discrepancy at higher $\\ell$ between observations and theories of the CMB polarization for models with a PMF [Panels (2c) and (2d) of Fig.2]. In our previous work \\cite{Yamazaki:2004vq,Yamazaki:2006bq} there was a problem from the strong degeneracy between $n_\\mathrm{B}$ and $B_\\lambda$. This degeneracy, however, is broken by the different effects of the PMF on the CMB power spectrum for lower and higher $\\ell$. The scalar-mode CMB temperature-anisotropy power-spectrum shape ($\\ell$-scaling for different $n_\\mathrm{B}$ at large scales - low $\\ell$) agrees with the results of \\cite{Kahniashvili:2007xe}, while the E-polarization power spectrum shape does not follow semi-analytical estimate in that paper. This difference is caused by the fact that we have included the effects of reionization \\cite{Dodelson:2003booka, Lewis02} which were neglected in \\cite{Kahniashvili:2007xe}. If the universe is re-ionized at $z_\\mathrm{re}$, CMB photons are scattered by electrons and the polarization is generated again. Since re-ionization results in a new scattering surface at relatively short distances from us at (and a relatively recent era), the viewing angle of the polarization from re-ionization becomes large. Thus, reionization causes some power to shift to lower multipoles \\cite{Kaplinghat:2002vt, Hu:2003gh}. We also point out that the observed power spectrum of temperature fluctuations in the CMB is likely to depend on frequency \\cite{Kuo:2006ya}. Such dependence is theoretically expected to originate from foreground effects such as the Sunyaev-Zel'dovich effect at higher multipoles. In contrast, the effects of a PMF are frequency-independent because the PMF affects the primary CMB as a background. Therefore, the correlation between the PMF and other foreground effects should be weak. Because of this, one should be able to eventually distinguish the PMF from foreground effects by using more than two observational data sets at different frequencies. In summary, we have found that we can constrain more precisely the power law index $n_\\mathrm{B}$ and amplitude $B_\\lambda$ of the PMF from all modes of CMB temperature and polarization anisotropies. The strong degeneracy of these parameters \\cite{Yamazaki:2004vq,Yamazaki:2006bq} is broken by the different effects of the PMF on the CMB power spectrum for lower and higher $\\ell$. The scalar mode from the PMF can be a main source for lower $\\ell$, while the vector mode can dominate for higher $\\ell$ in the CMB temperature anisotropies. Furthermore, these calculations suggest that it is possible to place a limit on the correlation parameters $s^\\mathrm{(X)}$ for large negative values of the spectral index. For example , $s^\\mathrm{(T)}, s^\\mathrm{(S)}< 0$ for $n_\\mathrm{B} = -2.9$ is ruled out from the effects of the TT mode on both the lowest and highest multipoles as shown in panels 2a and 2b of Figure 2. Such results may constrain models for the origin of the PMF, along with other PMF parameters." }, "0801/0801.2277.txt": { "abstract": "We present new spectroscopic and photometric data of the type Ibn supernovae 2006jc, 2000er and 2002ao. We discuss the general properties of this recently proposed supernova family, which also includes SN 1999cq. The early-time monitoring of SN 2000er traces the evolution of this class of objects during the first few days after the shock breakout. An overall similarity in the photometric and spectroscopic evolution is found among the members of this group, which would be unexpected if the energy in these core-collapse events was dominated by the interaction between supernova ejecta and circumstellar medium. Type Ibn supernovae appear to be rather normal type Ib/c supernova explosions which occur within a He-rich circumstellar environment. SNe Ibn are therefore likely produced by the explosion of Wolf-Rayet progenitors still embedded in the He-rich material lost by the star in recent mass-loss episodes, which resemble known luminous blue variable eruptions. The evolved Wolf-Rayet star could either result from the evolution of a very massive star or be the more evolved member of a massive binary system. We also suggest that there are a number of arguments in favour of a type Ibn classification for the historical SN 1885A (S-Andromedae), previously considered as an anomalous type Ia event with some resemblance to SN~1991bg. ", "introduction": "The most massive stars, i.e. those with initial masses larger than 30$M_\\odot$, are thought to end their lives with the core-collapse explosion of a Wolf-Rayet (WR) star. Before this stage, many of them undergo a short period of instability called the luminous blue variable (LBV) phase. This phase probably lasts of order 10$^{4}$-10$^{5}$ years \\citep{mae87,hum94,boh97,vink02,smi08}, during which they lose mass through recurrent mass-loss episodes. Most LBVs undergo fairly minor variability (typically below 1 magnitude) which is commonly called the S Doradus phase. But occasionally LBVs experience giant outbursts in which their luminosity rapidly increases and they can reach absolute visual magnitudes around $-14$ \\citep{dav97,mau05,vd06}, accompanied by the ejection of a large portion of their hydrogen envelope \\citep{smi6}. An overview of our current knowledge of LBVs can be found in \\citet{vge01}. At the time of explosion, which is expected to occur a few $\\times$ 10$^5$ years after the last major LBV outburst \\citep[see e.g.][]{heg03,eld04}, the mass of the WR star is expected to be of the order of 15-20$M_\\odot$. These stars may eventually explode as type Ib supernovae (SNe Ib) if they have lost the whole H envelope, or as type Ic supernovae (SNe Ic) if the stars are also stripped of their He mantles \\citep[for a comprehensive description of the different SN types see][]{fil97,tura07}. However, recent work \\citep{gal07,kot06,smi06} has suggested the possibility that some very massive stars may explode already {\\sl during} the LBV phase, producing luminous supernovae (SNe). Previously, a post-LBV channel was proposed by \\citet{sal02} to explain the observed properties of the type IIn SN~1997eg. In the proposed scenario, LBVs may eventually produce SNe which show high luminosity, blue colour, narrow H$\\alpha$ line in emission, and very slow spectro-photometric evolution. Their spectra are dominated by very narrow ($\\lesssim$1000 km s$^{-1}$), prominent H emission lines which sometimes show complex, multicomponent profiles. These objects would therefore belong to the so-called {\\sl type IIn} SN class \\citep{sch90}. The observed properties of these events are consistent with a scenario in which a significant fraction of the kinetic energy of the SN ejecta is converted into radiation via interaction with a dense, H-rich circumstellar medium (CSM). \\begin{figure} \\includegraphics[width=8.5cm]{figure1.ps} \\caption{Late R-band images of SN~2006jc. Liverpool Telescope image obtained on December 22, 2006 (left) and a very late VLT image (March 22, 2007; right). In this image, the SN is only marginally detectable and it is very close to a luminous star-forming region. North is down, East is to the left. \\label{fig1}} \\end{figure} A new clue for the death of the most massive stars has recently been proposed for the explosion of the peculiar SN 2006jc \\citep{ita06,pasto07,fol07,smi07,tom07,seppo07}. Extensive data sets spanning many wavelength regions have been also presented by \\citet{imm07b,anu07,elisa07,sak07}. \\cite{pasto07} showed the amazing discovery that a major outburst, reaching M$_R = -14.1$ (i.e. an absolute magnitude which is comparable to that of LBV eruptions), occurred only two years before the explosion of SN 2006jc, at exactly the same position as the SN and inferred that the two events must be physically related. %If the major outburst occurred in an LBV phase it would be surprising that %core-collapse happened so soon afterwords, at least from a theoretical stellar %evolution viewpoint. The physical connection between the two transients is supported by analysis of the SN spectra. These are indeed blue and dominated by relatively narrow ($\\sim$2200 km s$^{-1}$) He I lines in emission \\citep[hence the new classification as {\\sl type Ibn},][]{pasto07}, likely indicating a type Ib/c SN embedded within a dense, massive He-rich envelope. Unlike SNe IIn, the photometric evolution of this object is extremely rapid, while the spectra do not change as quickly as the luminosity. The best studied object of this type is SN~2006jc, but three additional events which appear to be spectroscopically rather similar have been found: SNe~1999cq, 2000er and 2002ao \\citep[discovered by][respectively]{mod99,cha00,mar02}. \\citet{fol07} first proposed that 2006jc-like events may constitute a distinct class of core-collapse SNe exploding in a He-rich CSM. SN~2005la \\citep{puk05} is also a somewhat similar SN, although with some unique properties. One might reasonably consider it as intermediate between a type Ibn and a type IIn event \\citep[see][]{pasto07b}. \\begin{figure*} \\includegraphics[width=8.4cm]{figure2a.ps} \\includegraphics[width=8.4cm]{figure2b.ps} \\caption{R-band images of SN 2000er (left) and SN 2002ao (right). They were obtained on November 26, 2000 using the ESO 3.6m Telescope in La Silla (Chile) and on February 7, 2002 using the IAC 80-cm, in La Palma (Canary Islands, Spain), respectively. \\label{fig2}} \\end{figure*} This article is the first of a series of three papers \\citep[together with][]{pasto07b,seppo07} in which we will discuss some peculiar, transitional events possibly produced by the explosions of very massive stars. In this work we will focus on the properties of the 4 objects which belong to the type Ibn SN family. Some of the data analysed in this paper have already been published \\citep{mat00,fol07,pasto07}, others have never been shown before. For SN 1999cq we do not present new data, but we consider only those of \\citet{mat00}. The afore mentioned SN~2005la, which has even more peculiar characteristics, will be analysed separately \\citep{pasto07b}. The paper is organised as follows: in Section \\ref{host} the properties of the host galaxies of the SN sample are introduced, including reddening and distance estimates. Spectroscopic and photometric observations are presented in Section \\ref{sp} and Section \\ref{lc}, respectively. The quasi-bolometric light curve of SN~2006jc is presented in Section \\ref{bolo}, while the parameters derived from the bolometric light curve modelling are discussed in Section \\ref{model}. In Section \\ref{prog} we examine two different plausible scenarios for the progenitors of SN~2006jc and similar events, while in Section \\ref{rate} we estimate the frequency of type Ibn events. Finally, a short summary is reported in Section \\ref{end}. ", "conclusions": "\\label{end} We have presented new data for three type Ibn SNe, i.e. SN~2000er, SN~2002ao and SN~2006jc. In particular, the early-time data of SN~2000er show, for the first time, the early spectroscopic behaviour of a type Ibn event. Early-time spectra of SN~2000er show relatively broad wings in the He I lines, with additional much narrower components. The detection of broad He I spectral features is unequivocal evidence for the presence of He also in the SN ejecta, not only in the CSM. Therefore, a residual envelope of He might be present in the progenitors of some SNe Ibn. Finally, the very late spectra of SN~2006jc still show strong narrow circumstellar He I lines, together with an emerging H$\\alpha$ likely produced in another CSM region. This data set allows us to highlight some key points concerning this new class of SNe, since we have now a better picture of the overall spectro-photometric evolution of SNe Ibn, which is essential to better constrain the characteristics of this class: i) SNe Ibn show a surprisingly high degree of homogeneity, which is unexpected if the ejecta were strongly interacting with surrounding CSM; ii) the modelling of the bolometric light curve of SN~2006jc (assuming no ejecta-CSM interaction) suggests that an amount of 0.2-0.4M$_\\odot$ of $^{56}$Ni was ejected. The mass of $^{56}$Ni could even be slightly higher for the brightest SNe Ibn (e.g. SN~1999cq). However, we cannot exclude that the interaction between SN ejecta and CSM may power the light curves of SNe Ibn to some degree, hence the $^{56}$Ni mass might be somewhat lower; iii) the observed data can be modelled without invoking extraordinarily massive ejecta; iv) their intrinsic rarity (being $\\sim$ 1$\\%$ of all core-collapse SNe) is probably due to the fact that they arise from a rather exotic progenitor scenario, consistent with either a post-LBV/early-WR channel of a very massive (60-100M$_\\odot$) star, or a binary (or even multiple) system formed by a normal LBV plus a core-collapsing WR companion." }, "0801/0801.0741_arXiv.txt": { "abstract": "% {Since 1999, we have been conducting a radial velocity survey of 179 K giants using the Coud\\'e Auxiliary Telescope at UCO/Lick observatory. At present $\\sim$20$-$100 measurements have been collected per star with a precision of 5 to 8 m s$^{-1}$. Of the stars monitored, 145 (80\\%) show radial velocity (RV) variations at a level $>$20 m s$^{-1}$, of which 43 exhibit significant periodicities.} {Our aim is to investigate possible mechanism(s) that cause these observed RV variations. We intend to test whether these variations are intrinsic in nature, or possibly induced by companions, or both. In addition, we aim to characterise the parameters of these companions.} {A relation between $\\log g$ and the amplitude of the RV variations is investigated for all stars in the sample. Furthermore, the hypothesis that all periodic RV variations are caused by companions is investigated by comparing their inferred orbital statistics with the statistics of companions around main sequence F, G, and K dwarfs.} {A strong relation is found between the amplitude of the RV variations and $\\log g$ in K giant stars, as suggested earlier by Hatzes \\& Cochran (1998). However, most of the stars exhibiting periodic variations are located above this relation. These RV variations can be split in a periodic component which is not correlated with $\\log g$ and a random residual part which \\textsl{does} correlate with $\\log g$. Compared to main-sequence dwarf stars, K giants frequently exhibit periodic RV variations. Interpreting these RV variations as being caused by companions, the orbital parameters are different from the companions orbiting dwarfs.} {Intrinsic mechanisms play an important role in producing RV variations in K giants stars, as suggested by their dependence on $\\log g$. However, it appears that periodic RV variations are \\textsl{additional} to these intrinsic variations, consistent with them being caused by companions. \\textbf{\\rm If indeed the majority of the periodic RV variations in K giants is interpreted as due to substellar companions, then massive planets are significantly more common around K giants than around F, G, K main-sequence stars.}} ", "introduction": "For more than a decade, radial velocity observations with accuracies of order m\\,s$^{-1}$ have been within reach (see for instance \\citet{marbut2000} and \\citet{queloz2001}). Even accuracies of less than 1 m\\,s$^{-1}$ \\citep{pepe2003} are possible now. With these observations, more than 200 sub-stellar companions have been discovered by measuring the reflex motions of their parent stars. Most of these sub-stellar companions have been detected around F, G and K main sequence stars, but detections around an A star \\citep{galland2006} and several subgiants (\\citet{johnson2006}, \\citet{johnson2007}) have also been reported recently. Moreover, 10 giant stars were reported to have sub-stellar companions ($\\iota$ Draconis (K2III) \\citet{frink2002}, HD104985 (G9III) \\citet{sato2003}, HD47526 (K1III) \\citet{setiawan2003}, HD13189 (K2II-III) \\citet{hatzes2005}, HD11977 (G5III) \\citet{setiawan2005}, Pollux (K0III) \\citet{hatzes2006}, \\citet{reffert2006}, 4UMa (K1III) \\citet{dollinger2007}, NGC2423 No3 and NGC4349 No127 \\citet{lovis2007}, and recently HD17092 (K0III) \\citet{niedzielski2007})\\footnote{For updated information on sub-stellar companions, see http://exoplanet.eu and http://exoplanets.org.}. In addition to searches for extra-solar companions, radial velocity observations prove to be very useful for detecting solar-like oscillations in stars with turbulent atmospheres, such as the dwarf $\\alpha$ Cen A \\citep[e.g.][]{bedding2006}, the subgiant Procyon \\citep[e.g.][]{eggenberger2004,martic2004} and the giant $\\epsilon$ Ophiuchi \\citep[e.g.][]{deridder2006}. With techniques for accurate radial velocity observations at hand, a survey was started in 1999 to verify whether K giants are stable enough to be used as astrometric reference stars for SIM/PlanetQuest (Space Interferometry Mission) \\citep{frink2001}. This survey contains 179 stars and uses the Coud\\'e Auxiliary Telescope (CAT) at University of California Observatories / Lick Observatory, in conjunction with the Hamilton Echelle Spectrograph. The survey has recently been expanded to about 380 giants and is still ongoing. For the analysis described in the present paper only data from the initial 179 stars are used. From this survey, companions have been announced for $\\iota$ Draconis \\citep{frink2002} and Pollux \\citep{reffert2006}. Stars with radial velocity variations of less than 20 m\\,s$^{-1}$ have been presented as stable stars by \\citet{hekker2006a}. In addition, some binaries discovered with this survey, as well as an extensive overview of the sample, will be presented in forthcoming papers. % As almost all of the stars show significant radial velocity variations, we investigate here which mechanism causes these variations. Non-periodic radial velocity variations, of the order of the investigated timescales, are most likely caused by some intrinsic mechanism, while the periodic variability can also be caused by companions. We also investigate the characteristics of these companions. In Sect. 2, the radial velocity observations are described in detail. In Sect. 3, the relation between the observed radial velocity amplitude and surface gravity is investigated. In Sect. 4, we explore the hypothesis that all periodic radial velocity variations are caused by sub-stellar companions, and we compare the inferred orbital parameters with those obtained for sub-stellar companions orbiting main sequence stars. Our conclusions are presented in Sect. 5. \\begin{figure} \\begin{minipage}{\\linewidth} \\begin{minipage}{\\linewidth} \\centering \\includegraphics[width=\\linewidth]{8321f1a.eps} \\end{minipage} \\begin{minipage}{\\linewidth} \\centering \\includegraphics[width=\\linewidth]{8321f1b-90.eps} \\end{minipage} \\end{minipage} \\caption{Radial velocity variations as a function of phase for a star (HIP34693) with a highly significant period (top), with its periodogram (bottom). The dashed line in the periodogram indicates the significance threshold.} \\label{vradphase1} \\end{figure} \\begin{figure} \\begin{minipage}{\\linewidth} \\begin{minipage}{\\linewidth} \\centering \\includegraphics[width=\\linewidth]{8321f2a.eps} \\end{minipage} \\begin{minipage}{\\linewidth} \\centering \\includegraphics[width=\\linewidth]{8321f2b-90.eps} \\end{minipage} \\end{minipage} \\caption{Radial velocity variations as a function of phase for a star (HIP7607) with a period close to the significance threshold (top), with its periodogram (bottom). The dashed line in the periodogram indicates the significance threshold.} \\label{vradphase2} \\end{figure} ", "conclusions": "The tight correlation we found between $\\log g$ and half of the peak-to-peak radial velocity variations seems to indicate that a large fraction of the observed radial velocity variations in our sample of K giants is induced by mechanism(s) intrinsic to the stars. We also present evidence that both intrinsic and extrinsic mechanisms play a role. The stars with a significant periodic signal are almost all located above the radial velocity amplitude vs. $\\log g$ relation, but when the periodic signal is removed, the residuals show the same trend as for the non-periodic stars. Furthermore, no correlation is present between the amplitude of the periodic signal and $\\log g$. Almost all of the lowest $\\log g$ stars show periodic variations. It may be possible that stars with such low surface gravity cannot be constant and that in these dilute atmospheres instabilities can occur very easily, and therefore may be periodic, but not extrinsic. Based on the evidence that extrinsic mechanism(s) play a role for K giant stars with periodic radial velocity variations we investigated the hypothesis that this periodic signal is caused by the reflex motion of sub-stellar companions orbiting these stars. We presented the characteristics of the orbital parameters of these companions and compared them with the known orbital parameters of sub-stellar companions orbiting F, G and K dwarfs. About 25$\\%$ of the stars in our sample have radial velocity variations with significant periodicity, and could possibly harbour a sub-stellar companion, while approximately only $8\\%$ of the 1330 F, G and K main sequence stars investigated by \\citet{marcy2005} have a sub-stellar companion. Recently \\citet{johnson2007} and \\citet{lovis2007} showed that the number of companion harbouring stars increases with mass. The giants in the present sample have typical masses between 1 and 4 M$_{\\sun}$ and are in general more massive than the main sequence stars investigated for companions. So, the high percentage is qualitatively in agreement with the results from the literature. Furthermore, \\citet{lovis2007} suggest that more massive stars form more massive planetary systems than lower mass stars. Figs~\\ref{msini} and \\ref{msinilow} show that we find, in general, more massive companions around the more massive K giants than are present around F, G and K dwarfs. The high percentage of more massive companions around the more massive K giant stars would also be compatible with the core accretion model. This model predicts very few giant planets, but a relatively large number of planets with the mass of Neptune or smaller around M dwarfs \\citep{laughlin2004,ida2005}. This is mainly the result of a much reduced surface density of the disk and the resulting shorter disk evolution timescales compared to those for more massive stars, implying that planet properties vary with the mass of their host stars. In particular, one would expect more sub-stellar companions with higher masses in our giant sample, as is indeed the case if we assume that the companion hypothesis is correct. The mean metallicity of companion-hosting K giants would be similar to the mean metallicity of the total sample. This is in contrast with the correlation between companion occurrence and metallicity present in F, G and K dwarfs \\citep[e.g.][]{fischer2005}. So far, several groups have investigated the correlation between companion occurrence and metallicity for giants with different results. \\citet{sadakane2005} and \\citet{pasquini2007} agree that companion-ahosting giants are on average not metal-rich, while \\citet{hekker2007} find that giants with announced companions have higher metallicities than their total sample of giants. A detailed discussion about these different results is presented by \\citet{hekker2007}. They conclude that the samples on which the results are based are slightly different and the more metal-rich stars used in their study are lacking in the study by \\citet{pasquini2007}. Furthermore, there is a difference in zero-point correction for the metallicities of announced companion-hosting stars from different surveys. All in all, these inferences are based on small-number statistics and all results have to be taken with caution. The larger semi-major axis and long periods of the inferred companions orbiting K giant stars compared to companions orbiting dwarf stars are most likely due to the extended atmospheres of K giants. For the eccentricity no significant difference is found in the distribution between companions orbiting dwarfs or giants. Nevertheless, the high number of inferred companions around giants with eccentricities $<$ 0.3 is striking. One could suspect that companion orbits circularise over time and that the companions in circular orbits are older than the eccentric ones, but there is no evidence for this hypothesis. In principle, nearly sinusoidal radial velocity variations could also be caused by pulsations or spots. Although the periods of the radial velocity variations could well be the rotational periods of the stars, the presence of prominent spots is not very likely. In that case one would also expect photometric variations with periods correlated with the radial velocity variations. From the Hipparcos photometry \\citep{esa1997} such correlations were not found. In order to distinguish with certainty between companions and pulsations as the cause of the observed radial velocity variations, one needs to perform a spectral line profile analysis. A technique for doing this with very high-resolution spectra ($R \\geq$ 100\\,000) will be presented separately (Hekker et al., in preparation) because the amount of such data at hand today is insufficient to add significantly to the conclusions of this paper. \\begin{acknowledgement} We thank Debra Fischer, Geoff Marcy and Paul Butler for useful discussions and the development of the instrumentation and software for the determination of the radial velocities at Lick Observatory. In addition, we thank the entire staff at Lick Observatory for their excellent support. Finally, we would like to thank the anonymous referee for valuable comments and suggestions. \\end{acknowledgement}" }, "0801/0801.0607_arXiv.txt": { "abstract": "We present the first attempt to analytically study the nonlinear matter power spectrum for a mixed dark matter (cold dark matter plus neutrinos of total mass $\\sim 0.1$eV) model based on cosmological perturbation theory. The suppression in the power spectrum amplitudes due to massive neutrinos is, compared to the linear regime, enhanced in the weakly nonlinear regime where standard linear theory ceases to be accurate. We demonstrate that, thanks to this enhanced effect and the gain in the range of wavenumbers to which the PT prediction is applicable, the use of such a nonlinear model may enable a precision of $\\sigma(m_{\\nu, {\\rm tot}})\\sim 0.07$eV in constraining the total neutrino mass for the planned galaxy redshift survey, a factor of 2 improvement compared to the linear regime. ", "introduction": " ", "conclusions": "" }, "0801/0801.2602_arXiv.txt": { "abstract": "Pulsar winds shocked in the ambient medium produce spectacular nebulae observable from the radio through $\\gamma$-rays. The shape and the spectrum of a pulsar wind nebula (PWN) depend on the angular distribution, magnetization and energy spectrum of the wind streaming from the pulsar magnetosphere, as well as on the pulsar velocity and the properties of the ambient medium. The advent of {\\sl Chandra}, with its unprecedented angular resolution and high sensitivity, has allowed us not only to detect many new PWNe, but also study their spatial and spectral structure and dynamics, which has significantly advanced our understanding of these objects. Here we overview recent observational results on PWNe, with emphasis on {\\sl Chandra} observations. ", "introduction": "It is generally accepted that all active pulsars lose their spin energy and angular momentum via relativistic winds comprised of relativistic particles and electromagnetic field. Because the relativistic bulk velocity of the wind leaving the pulsar magnetosphere is obviously supersonic with respect to the ambient medium, such a wind produces a {\\em termination shock} (TS) at a distance $R_{\\rm TS}$ from the pulsar where the bulk wind pressure, $P_w\\sim\\edot/(4\\pi c R_{\\rm TS}^2)$, is equal to the ambient pressure $P_{\\rm amb}$. The TS radius can be estimated as \\be R_{\\rm TS} \\sim \\left(\\edot/4\\pi c P_{\\rm amb}\\right)^{1/2} \\sim 0.05\\, \\edot_{36}^{1/2} P_{\\rm amb,-10}^{-1/2}\\,\\,\\,\\,{\\rm pc}, \\ee where $\\edot = 10^{36}\\edot_{36}$ ergs s$^{-1}$ is the pulsar's spin-down power, and $P_{\\rm amb}=10^{-10} P_{\\rm amb,-10}$ dyn cm$^{-2}$. At the TS, the pulsar wind is being ``thermalized'', and the downstream bulk flow speed becomes subrelativistic \\cite{1984ApJ...283..710K}. As the relativistic particles of the shocked wind move in the magnetic field and ambient radiation field, they emit synchrotron and inverse Compton (IC) radition, which we observe as a {\\em pulsar wind nebula} (PWN). Since the wind is a universal property of any active pulsar, we expect that {\\em all pulsars must be accompanied by PWNe}. Studying PWNe tells us about the properties of pulsar winds and their parent pulsars, the properties of the ambient medium, and the mechanisms of wind-medium interaction. The energies of the synchrotron and IC photons span the range from the radio to TeV $\\gamma$-rays. The mere detection of a PWN in some energy band indicates the emission mechanism and the electron energies involved. For instance, detecting a PWN in the X-ray band, where the synchrotron emission dominates, implies that the wind particles have been accelerated up to $\\gtrsim 100$ TeV energies, either at the TS or on the way to the TS (note that particles with such energies cannot leave the pulsar magnetosphere because of radiative losses), and that the same particles should produce IC emission in the TeV energy range. Useful information on the pulsar wind and its interaction with the medium is provided by the PWN morphology, which depends on the wind {\\em outflow geometry} and the direction of {\\em pulsar velocity}. If we assume an {\\em isotropic outflow} from a very slowly moving pulsar, the shocked wind is confined between the TS and {\\em contact discontinuity} (CD) spheres, while the shocked ambient medium fills in the space between the CD and the spherical {\\em forward shock} (FS). If the pulsar moves with a supersonic velocity, $V_p \\gg c_s$, then the TS, CD and FS surfaces acquire a characteristic convex shapes ahead of the moving pulsar but can exhibit rather different shapes behind it. In particular, the TS acquires a bullet-like shape, with the distance $R_{\\rm TS,h}\\sim (\\edot/4\\pi c P_{\\rm ram})^{-1/2}\\sim 0.04 \\edot_{36}^{1/2}n^{-1/2}(V_p/100\\,{\\rm km\\,s}^{-1})^{-1}$ pc, between the bullet's head ({\\em bowshock}) and the pulsar (cf.\\ eq.\\ 1), where $P_{\\rm ram} = \\rho V_p^2 = 1.67\\times 10^{-10} n (V_p/100\\, {\\rm km\\,s}^{-1})^2$ dyn cm$^{-2}$ is the ram pressure. The numerical simulations \\cite{2003A&A...404..939V, 2005A&A...434..189B} suggest that the shocked wind is channeled into the {\\em tail} behind the TS bullet, confined by the nearly cylindrical CD surface, where it flows with a mildly relativistic velocity. The shocked ISM matter is also stretched along the pulsar trajectory, and it can be seen in spectral lines (e.g., H$_\\alpha$) from the atoms excited at the FS (e.g., \\cite{2002ApJ...575..407C}). \\begin{figure}[h] \\label{p-pdot} \\hspace{-0.3in} \\includegraphics[height=3.3in,width=3.1in,angle=90]{P_Pdot.pdf} \\caption{$P$-$\\dot{P}$ diagram for the pulsars in the ATNF catalog \\cite{2005AJ....129.1993M}. Pulsars with known X-ray PWNe are marked by circles. } \\end{figure} \\begin{figure} \\centering \\includegraphics[width=6.5in,angle=0]{tori3_small.pdf} \\caption{X-ray images of PWNe with toroidal components. The image numbers correspond to Tables 1--2. } \\end{figure} We know from observations of young, subsonically moving pulsars (such as the Crab pulsar), that the PWN is not spherical even in this case, but it rather looks like a torus, sometimes with one or two jets along the pulsar's spin axis. This means that the pre-shock wind is {\\em not isotropic}, but it outflows preferentially in the equatorial plane of the rotating pulsar. Models of such {\\em torus-jet} PWNe have been simulated in \\cite{2004MNRAS.349..779K} and \\cite{2006A&A...453..621D} (see also Bucciantini, this proceedings). We are not aware of relativistic MHD models for PWNe created by supersonically moving pulsars with {\\em anisotropic} outflows, but we expect that the compact PWN morphology near the pulsar may significantly depend on the orientation of the spin axis with respect to the direction of pulsar's motion, while the distant PWN tail and the FS should not be strongly affected by the wind anisotropy. As the PWN appearance critically depends on the pulsar's Mach number, ${\\mathcal M}=V_p/c_s$, and $c_s$ in a hot SNR interior is much higher than that in the normal ISM, we expect that young, powerful pulsars, which have not left their host SNRs, generate torus-jet PWNe, while older pulsars would show bowshock-tail PWNe \\cite{2003A&A...397..913V}. Only a handful of PWNe had been detected before the launch of \\chan, mostly in radio and H$_\\alpha$ observations \\cite{2002ApJ...575..407C}, while X-ray observations of PWNe had been hindered by low angular resolution of X-ray instruments. \\chan, with its unprecedented $\\approx 0.5''$ resolution, has allowed us to reveal the fine structure of the previously known X-ray PWNe, discover many new PWNe, and separate the pulsar and PWN emission in many cases. Some interesting results on X-ray PWNe have also been obtained with {\\sl XMM-Newton}, which lacks the high angular resolution of \\chan\\ but is more sensitive and has a larger field of view. Many exciting results from the \\chan\\ and \\xmm\\ observations of PWNe have been reported in numerous publications, including two reviews \\cite{2006csxs.book..279K, 2006ARA&A..44...17G}. Here, we present an up-to-date overview of X-ray observations of PWNe, including a gallery of spectacular PWN images and a current catalog of these objects, and discuss correlations between various PWN properties as well as the pulsar-PWN correlations. ", "conclusions": "The high angular resolutions and sensitivity of \\chan\\ have allowed us to detect many X-ray PWNe and study their structure and spectra. Our current understanding of X-ray PWNe can be briefly summarized as follows. \\begin{itemize} \\item Most of the detected PWNe are associated with young, powerful pulsars (partly due to selection effect), but some old PWNe have also been detected (e.g., the tail of PSR B1929+10). \\item The observed PWN morphologies can be crudely classified into the torus-jet and bowshock-tail types, corresponding to sub- and supersonic pulsar motion, respectively. However, the classification is often uncertain, and many morphological features (e.g., in the Vela and Geminga PWNe) remain to be understood. \\item PWNe radiate up to a few percent of the pulsar spin-down power in X-rays. The X-ray PWN efficiency generally grows with $\\edot$, but the $\\etapwn$-$\\edot$ dependence shows a very large scatter, including some very dim PWNe, $\\etapwn < 10^{-5}$. \\item The X-ray luminosities of the detected PWNe and their parent pulsars are of the same order of magnitude, with PWNe being, on average, somewhat more luminous. \\item The photon indices of the PWN spectra are in the range $1\\lesssim \\Gamma \\lesssim 2$. More X-ray efficient PWNe apparently show softer spectra, but this conclusion should be checked in deeper observations of well-resolved PWNe to mitigate the effects of synchrotron cooling. \\item Spectral maps of the Crab and Vela PWN show strong correlation between the spectral hardness and morphological features. The Crab's spectra immediately downstream of TS are considerably softer than those of the Vela. \\item X-ray observations of pulsars/PWNe in the vicinity of extended TeV sources show faint extensions of compact PWNe toward the offset centers of the TeV emission, supporting the interpretation of these TeV sources as ``crushed plerions''. \\item Parsec-scale X-ray tails found behind many pulsars represent ram-pressure collimated, high-velocity flows of relativistic particles, resembling pulsar jets in their properties. \\end{itemize} Notwithstanding the considerable progress in observations and modeling of PWNe, there still remains a nubmer of important problems to solve. Here we mention a few of them. \\begin{itemize} \\item What is the origin of the diverse, often very complex, PWN morphologies? For instance, what is the nature of the outer arc, inner jets, and bar in the Vela PWN, the loops in the 3C\\,58 PWN, the shell and axial tail in the Geminga PWN? \\item Where and how the X-ray emitting particles are accelerated? Does the acceleration occur in the preshock wind or at the TS? Are the winds composed of electrons/positrons or they have some nucleonic component? \\item What are the physical parameters that determine the X-ray PWN efficiency, in addition to $\\edot$? Why don't we see any PWN around some pulsars? Is it because of some special properties of the pulsar wind or the ambient medium? \\item Why are the particle spectra so hard in some PWNe (e.g., Vela) and much softer in others (e.g., Crab)? Is such a difference caused by different properties of the parent pulsars' winds (e.g., magnetization), or different efficiencies of particle acceleration between the magnetosphere and the TS? What is the reason for the apparent efficiency-hardness correlation? \\item Can all the Galactic extended TeV sources with luminous pulsars nearby be interpreted as relic TeV plerions or there are indeed ``dark accelerators''? Are these sources indeed ``crushed PWNe'' formed by the passage of the inverse SNR shock? Can the huge sizes of the TeV plerions and the large offsets from the pulsars be reconciled with the ``crushed PWN'' scenario? Is the TeV radiation due to the IC scattering or the pion decay? If the former, are the seed photons supplied by the CMB or IR radiation from nearby star-forming regions or dust clouds? \\item What are the flow velocities and magnetic fields in the long pulsar tails? Which role does the magnetic field play in the tail collimation and what is the magnetic field topology ? How do these collimated flows decelerate and cool? \\end{itemize} To answer these questions, new observations are badly needed, both in X-rays and other wavelengths, as well as theoretical work and numerical modeling. Particularly important would be to take full advantage of the outstanding \\chan\\ capabilities as long as it is alive because no X-ray observatory with such high angular resolution is expected in the foreseeable future. \\begin{theacknowledgments} We thank Koji Mori for providing the {\\sl ASCA} images of Vela X. This work was partially supported by Chandra award AR5-606X. \\end{theacknowledgments}" }, "0801/0801.2328_arXiv.txt": { "abstract": "Shell galaxies are widely considered the debris of recent accretion/merging episodes. Their high frequency in low density environment suggests that such episodes could be among the driver of the early-type galaxy secular evolution. We present far and near UV (FUV and NUV respectively hereafter) {\\it GALEX} photometric properties of a sample of shell galaxies. ", "introduction": "In a hierarchical evolutionary scenario, galaxies experience accretion/merging events during their lifetime. While early-type galaxies in nearby clusters appear (homogeneously) old, the field early-type galaxy population seems to contain genuinely, recently {\\it rejuvenated} objects \\citep[see e.g.][]{Clemens06}. Early-type galaxies showing fine structure, like shells, occupy a special position since they are believed to fill the gap between ongoing mergers and normal elliptical galaxies. The UV emission is crucial to test whether these galaxies do host ongoing/recent star formation processes and study their distribution across the galaxy. We present new {\\it GALEX} observations of three shell galaxies NGC 1210, MGC -05-07-1 (GI04-0030-0059 PI D. Bettoni) and NGC 5329 (from archive) in addition to those analyzed in \\citet{Rampazzo07} which we use as baseline for our preliminary conclusions. ", "conclusions": "Table 1 summarizes the journal and the basic results of the {\\it GALEX} observations. \\begin{table*} \\scriptsize \\caption{Journal of the {\\it GALEX} observations} \\begin{tabular}{lcclllll} \\hline \\multicolumn{1}{l}{Name}& \\multicolumn{1}{c}{NUV} & \\multicolumn{1}{c}{FUV} & \\multicolumn{1}{l}{P.I.}& \\multicolumn{1}{c}{m$_{NUV}^{tot}$}& \\multicolumn{1}{c}{m$_{FUV}^{tot}$} & \\multicolumn{1}{c}{FUV-NUV}\\\\ \\multicolumn{1}{c}{}& \\multicolumn{1}{c}{exposure [sec]} & \\multicolumn{1}{c}{exposure [sec]} & \\multicolumn{1}{c}{}& \\multicolumn{1}{c}{}& \\multicolumn{1}{l}{}& \\multicolumn{1}{l}{}\\\\ \\hline MCG-05-07-1 &1510 & 1531 & D. Bettoni & 18.67$\\pm$0.07 & 19.76$\\pm$0.07 & 1.11$\\pm$0.10 \\\\ NGC 1210 &1558 & 1608 & D. Bettoni &17.14$\\pm$0.02&20.08$\\pm$0.07&2.95$\\pm$0.07 \\\\ NGC5329 &3889 &2666 & MIS&18.20$\\pm$0.02&20.20$\\pm$0.04&2.02$\\pm$0.05 \\\\ \\hline \\end{tabular} \\label{table1} \\medskip{FUV - NUV AB magnitudes have been corrected for Galactic extinction.} \\end{table*} Smoothed images and 2D colour maps are shown in Figure 1. FUV emission in NGC 5329 is present only in the central part of the galaxy. In NGC 1210 and MGC-05-07-01 the FUV is quite strong both in the polar ring of the former galaxy and in the debris systems, residual of the accretion events of both galaxies. The (FUV-NUV) colour in the north tail of NGC 1210 ($\\sim$0.31, $\\sim$0.34, $\\sim$0.96), in the polar ring ($\\sim$0.36, $\\sim$0.73) and in the nucleus ($\\sim$1.11) of MGC-05-07-01 are quite blue. \\citet{Neff05} show that tail/bridges produced by interaction may have similar colour indicative of a recent star formation (200-300 Myr). The age of 1-3 Gyr estimated by Whitmore et al. (1987), responsible for the present structure of MCG-05-07-01, is consistent with age indication coming from (UV - optical) colours. \\begin{figure} \\centerline{\\psfig{figure=marino1.ps,width=2.8cm}\\psfig{figure=marino2.ps,width=2.8cm}}% \\centerline{\\psfig{figure=marino3.ps,width=2.8cm}\\psfig{figure=marino4.ps,width=2.8cm}}% \\centerline{\\psfig{figure=marino5.ps,width=2.8cm}\\psfig{figure=marino6.ps,width=2.8cm}}% \\caption{{\\it GALEX} FUV (left panels), NUV (middle panels) images (5\\arcmin $\\times$ 5\\arcmin) and (FUV-NUV) colour maps (right panels) of MCG-05-07-1 (top row), NGC 1210 (mid row) and NGC 5329 (bottom row). FUV and NUV images are smoothed using {\\it Asmooth} \\citep{Ebeling06}.} \\label{fig1} \\end{figure}" }, "0801/0801.3432_arXiv.txt": { "abstract": "Difference imaging provides a new way to discover gravitationally lensed quasars because few non-lensed sources will show spatially extended, time variable flux. We test the method on lens candidates in the Sloan Digital Sky Survey (SDSS) Supernova Survey region from the SDSS Quasar Lens Search (SQLS) and their surrounding fields. Starting from 20768 sources, including 49 SDSS quasars and 36 candidate lenses/lensed images, we find that 21 sources including 15 SDSS QSOs and 7 candidate lenses/lensed images are non-periodic variable sources. We can measure the spatial structure of the variable flux for 18 of these sources and identify only one as a non-point source. This source does not display the compelling spatial structure of the variable flux of known lensed quasars, so we reject it as a lens candidate. None of the lens candidates from the SQLS survive our cuts. Given our effective survey area of order 0.71 square degrees, this indicates a false positive rate of order one per square degree for the method. The fraction of quasars not found to be variable and the false positive rate should both fall if we analyze the full, later data releases for the SDSS fields. While application of the method to the SDSS is limited by the resolution, depth, and sampling of the survey, several future surveys such as Pan-STARRS, LSST, and SNAP will avoid these limitations. ", "introduction": "Gravitational lensing has many applications, from exoplanet searches to large scale structure (see the reviews by Kochanek, Schneider, and Wambsganss in the Saas Fe lectures, 2006). Galaxy scale lenses can be used to study the dark matter mass profile of the lens galaxy \\citep[e.g.][]{Kochanek91,Rusin05,Koopmans06,Jiang07}. When the background source is a quasar, microlensing by individual stars in the lens galaxy can be used to probe the structure of the quasar's accretion disk \\citep{Poindexter07,Morgan07} and broad line regions \\citep{Eigenbrod07}. Galaxy scale lenses can also constrain cosmological parameters \\citep{Refsdal64,Oguri07}. Unfortunately, most applications require large samples of gravitational lenses to be competitive with other methods, while fewer than a hundred lensed quasars are known. Moreover, these lenses were discovered using different methods with different biases, which is especially problematic if homogeneous samples with well-understood selection functions are needed \\citep[see][]{Kochanek04}. The known lenses were found by their morphological structure or the presence of higher redshift features in the spectrum of a lower redshift galaxy. Morphological surveys examine optical or radio images of quasars for evidence that they are lensed. This works best for point-like sources like optical quasars, and flat spectrum radio sources. The two largest searches for lensed AGN are the Cosmic Lens All-Sky Survey of flat spectrum radio sources \\citep{Myers03,Browne03} and the SDSS Quasar Lens Search \\citep[SQLS;][]{Oguri06}. The radio samples are limited by the number of sufficiently bright radio sources and the difficulties in obtaining source redshifts. The optical quasar samples suffer from confusion from stars and galaxies and the effects of color changes, either from the starlight or dust in the lenses, on sample selection. The spectroscopic method made several serendipitous discoveries of lensed quasars, such as Q2237+0305 \\citep{Huchra85} and SDSSJ090334.92+502819.2 \\citep{Johnston03}. Following theoretical investigations \\citep[e.g.][]{Kochanek92,Miralda92,Mortlock00,Mortlock01}, the SLACS survey \\citep{Bolton04,Bolton06} used the spectroscopic method on massive early-type galaxies in the SDSS to identify $\\sim$50 candidate lensed starforming galaxies, and confirmed 19 of 29 using Hubble Space Telescope images. The spectroscopic method is limited by its low yield (about 1 lens candidate per 1000 luminous red galaxy spectra) and biases in its mass range (due to the size of the spectroscopic aperture). Moreover, by selecting targets based on the properties of the lens galaxy, these lenses are mainly useful for studying the lens galaxies rather than for cosmology \\citep[see][]{Kochanek04}. \\citet{Kochanek06} proposed a new method to find lensed quasars based on difference imaging. Difference imaging, also known as image subtraction, is a way to measure the variable intensity in a region of sky \\citep{Tomaney96,Alard98,Alard00}. It has been used to search for a broad range of variable sources, such as planets transiting stars \\citep[e.g.][]{Hartman04}, microlenses \\citep[e.g.][]{Alcock99,Wozniak00}, and supernovae \\citep[e.g. the Sloan Digital Sky Survey II Supernova Survey,][]{Sako07}. In difference imaging, a reference image is made by averaging a set of the best images for a field. Then, for each epoch of observation, a difference image is created by subtracting the reference image convolved to the PSF and flux scale of that epoch. If a source is variable, there will be a flux residual in the difference image corresponding to the variability of the source. In difference images, lensed quasars are recognizable because they consist of multiple variable images that are close together. Since most quasars are variable, with 60\\% having $\\sigma_g \\ge 0.05$ variability over two years \\citep[e.g.][]{Sesar07}, each of these images is variable. The level of variability is then enhanced by microlensing of the quasar images by the stars in the lens galaxy \\citep[see the review by][]{Wambsganss06}. Thus, lensed quasars look like compact clusters of variable sources or extended variable sources, which allows us to easily search for lensed quasars in difference images. The number of false positives from pairs of variable stars, pairs of quasars (related or not), variable star-quasar blends, and supernovae-AGN pairs is expected to be low \\citep{Kochanek06}. Candidates found this way can be confirmed by light curve analysis, since each image will have the same intrinsic variability, but with time delays between the images and some additional uncorrelated variability from microlensing by the stars of the lens galaxy \\citep[e.g.][]{Pindor05}. Difference imaging can also resolve variable lensed sources blended with non-variable sources, such as the lens galaxies. This is important because, as we search for fainter lensed quasars, contamination from the lens galaxy becomes a steadily greater problem. Tests of image subtraction on the lens Q2237+0305 show that the method can find quasars even when they are buried by the flux of an extraordinarily bright foreground galaxy \\citep{Kochanek06}. The forthcoming, large scale synoptic surveys, like LSST \\citep{Tyson02} and Pan-STARRS \\citep{Kaiser04}, will be ideal for this method, since they will cover large areas of the sky, sample variable sources frequently, and have deep magnitude limits. In the meantime, the SDSS II Supernova Survey \\citep{Frieman07,Sako07} is the best survey currently available for our goals. Intended to find supernovae for dark energy studies, the Supernova Survey repeatedly images SDSS Stripe 82, a 2.5\\degr~wide swath along the celestial equator, covering 300 square degrees and stretching across the Southern Galactic Cap from right ascension 300\\degr~to right ascension 60\\degr~\\citep{Frieman07}. Ten to twenty public epochs were available for a given field when we started this project, with the number rising to about forty in the most recent releases \\citep{Adelman-McCarthy07}. We applied the method outlined in \\citet{Kochanek06} to the SDSS Supernova Survey fields that contained 26 previously identified candidates for gravitational lenses, with 39 total components, from the initial SQLS candidate list. These include 15 candidates with 24 components in the main statistical sample \\citep{Inada07}, and 11 additional candidate systems, with 15 images, that were outside the final selection criteria. These candidates are listed in Table~\\ref{table:InadaLenses}. We also checked to see if any pair of lens images in a candidate lens appeared on the difference images, since the images can be well-separated. The lens candidates had already been rejected for other reasons, such as different spectral energy distributions for the postulated lens images, or the lack of a lens galaxy. Unfortunately, there are no confirmed lensed quasars in the Supernova Survey region. We had two goals. First, to use the variability method as an independent check of the SDSS candidates. Second, to get a sense of the false positive rate from any other variable sources in the fields. We outline the method, our approach to source selection, and the results when applied to targets in \\S 2. We discuss the variable sources in the lens candidate fields in \\S 3, including quasars (\\S 3.2) and an extended variable source (\\S 3.3). Finally, we conclude in \\S 4 by discussing prospects for finding gravitational lenses in future surveys with difference imaging. ", "conclusions": "We searched for sources with the spatially extended variability characteristic of gravitationally lensed quasars by applying difference imaging to the Sloan Supernova Survey fields of the SQLS lens candidates and their surroundings. We found one source, SDSSJ213245.25+000146.5, that passed basic criteria for variability, non-periodic variability, and which appeared extended on its ``absdiff'' image. However, it did not show the compelling structure seen in difference images of known lensed quasars \\citep{Kochanek06} -- it only barely passed the effective radius criterion, suggesting it is not likely to be a true lensed quasar. Furthermore, it was not among the SQLS lens candidates in the SDSS Supernova Survey region. Our criteria successfully rejected the SQLS candidates, as had \\citet{Inada07} using other criteria. If used with other lens search techniques, like color or morphological criteria \\citep{Oguri06}, difference imaging could substantially reduce the number of false positives, although it would only work for quasars that are variable during the observation period. Only a third of the quasars detected by ISIS passed our quasar variability tests and could be analyzed for spatial structure in the difference images. More generally, as a variability survey for lensed quasars covering 0.71 square degrees, we successfully identified one third of the SDSS quasars in the field and had no false positive detections of lenses. In many ways, our experiment was limited by the data. First, there are no known lensed quasars in Stripe 82, so we could evaluate no examples of success or set any limits on the role of false negatives. Furthermore, when we carried out this experiment, the Supernova Survey had few public ($\\la 25$) epochs available. Since some epochs will be dropped because of bad seeing, or because image subtraction failed, there are some fields where there were too few enough epochs to extract light curve statistics. Similarly, the released epochs did not cover a large enough time interval. In some cases (e.g. SDSSJ213245.25+000146.5 in Figure~\\ref{fig:LCs}), all the non-rejected epochs spanned only a few months. The performance of our approach would improve markedly with survey data spanning several years and with many more epochs. Additionally, the relatively poor resolution of the SDSS survey is not ideal for identifying quasar lenses. This is also true for color and morphological techniques, such as in \\citet{Inada07}, where candidate lenses with image separations of less than 1\\farcs0 were rejected. With better resolution and more epochs, false positives in the method would decrease. If we examined every SDSS quasar in the supernova region, we would expect to find of order 1000 were variable with $\\sim 50$ false positives if we assume no improvement in the false positive rate with the addition of more epochs. In reality, we should have a marked reduction in the rate because we could focus on epochs with better seeing and add in the variability information from the ugiz bands as well. While these shortcomings will be partly solved by the full SDSS Supernova Survey data set, the real future for the method lies with ground-based surveys like LSST \\citep{Tyson02} and Pan-STARRS \\citep{Kaiser04} and proposed surveys from space like the SuperNova Acceleration Probe \\citep[SNAP,][]{Aldering02}. Pan-STARRS and LSST would repeatedly survey very large areas ($10^3$--$10^4$ rather than $10^2$ square degrees) with improved resolution ($0\\farcs5$--$1\\farcs$ rather than $>1\\farcs0$) and depth ($\\simeq 24$~mag rather than $\\simeq 22$~mag). \\citet{Kochanek06} estimated that LSST can discover $\\sim 10^3$ lensed quasars with V$<23$. The space based surveys would cover far less area (15~square degree for SNAP) but with vastly improved resolution ($\\simeq 0\\farcs05$) and depth ($\\simeq 28$~mag). Since few lenses have separations this small (1\\farcs5 is the expected median, e.g. \\citet{Mitchell05}), any lens identified by variability is easily confirmed by its morphology. More important for the space missions is that their great depth allows searches for other lensed, time variable sources such as the supernovae themselves." }, "0801/0801.3268_arXiv.txt": { "abstract": "We apply the K-correction to the black hole LMXB GX 339-4 which implies $M_{X}\\geq 6 M_{\\odot}$ by only assuming that the companion is more massive than $\\sim 0.17M_{\\odot}$, the lower limit allowed by applying a 'stripped-giant' model. This evolutionary model successfully reproduces the observed properties of the system. We obtain a maximum mass for the companion of $M_2 \\leq 1.1 M_{\\odot} $ and an upper limit to the mass ratio of $q(=M_2/M_{X})\\leq0.125$. The high X-ray activity displayed by the source suggests a relatively large mass transfer rate which, according to the model, results in $M_2 \\ga 0.3M_{\\odot}$ and $M_{X} \\ga 7M_{\\odot}$. We have also applied this scenario to the black hole binary XTE J1550-564, which has a similar orbital period but the donor is detected spectroscopically. The model successfully reproduces the observed stellar parameters. ", "introduction": "\\label{introduction} Low Mass X-ray Binaries (LMXB) are interacting binaries harbouring a neutron star (NS) or a black-hole (BH) which accretes matter from a low mass star. A subgroup, the so-called X-ray transients (SXTs), provide a unique opportunity to set dynamical constraints to the the masses of both NSs and BHs because the spectrum of the companion dominates during the quiescent (X-ray off) states. In particular, we can empirically determine the mass function of the system which is usually expressed as: \\begin{equation} \\label{mf} f(M)=\\frac{PK_2^3}{2 \\pi G}=\\frac{M_X\\sin{i}^3}{(1+q)^2} \\end{equation} where $P$ is the orbital period, $K_2$ the semi-amplitude of the velocity curve of companion star, $M_X$ the mass of the compact object, $i$ is the orbital inclination of the systems and $q=\\frac{M_2}{M_X}$ the mass ratio. Therefore, it is possible to establish a secure lower limit to the mass of the compact object by only measuring $P$ and $K_2$. Although this method has successfully been applied to $\\sim 20$ BH X-ray binaries (see \\citealt{C2006} for details), they only represent the 'peak of the ice-berg' of a large population of $10^8-10^9$ stellar-mass BH present in our galaxy (e. g. \\citealt{BB94}). Most quiescent SXTs are optically faint, with $R\\geq 20$, and hence dynamical constraints are usually affected by large errors. Setting new dynamical constraints and refining the existing mass solutions are the only ways to get new insights into the fundamental properties of stellar-mass BHs.\\\\ GX 339-4 (V821 Ara) was discovered by OSO 7 in 1972 (\\citealt{Ma73}). Since then, several X-ray outburst (e. g. 4 in the last decade) have been observed, and all the X-ray states have been detected in this system (See \\citealt{Mvdk97} for the intermediate state and \\citealt{Mi91} for the Very High state). Moreover, \\cite{Cor00} detected a compact jet emission from this system, showing that it is a microquasar. GX 339-4 was early proposed as a BH candidate based on its X-ray properties (\\citealt{S79}) but the spectral features of the companion star could not be detected even during 'X-ray off' states (\\citealt{Tariq01}), preventing a dynamical confirmation (i. e. $M_X > 3M_{\\odot}$). Only during the 2002 outburst \\cite{hynes03} reported the first detection of the donor star thanks to the discovery of NIII/CIII Bowen emission lines arising from the irradiated companion star. The lines are very sharp and swing with a velocity semi-amplitude of $K_{em}$=$317 \\pm 10 $ km s$^{-1}$. These authors also reported an orbital period of $P=1.7557\\pm0.0004$ days which was later confirmed by \\cite{lyc06}. % The combination of both, radial velocity of the companion and orbital period yields $f(M)=5.8 \\pm 0.5 M_{\\odot}$ (and hence $M_X \\geq 5.3M_{\\odot}$), which represents the first dynamical proof for a BH in GX 339-4. Note that narrow high-excitation emission lines originating from the companion star have also been detected in many others LMXBs (e.g. \\citealt{C04}) and demonstrate that this is a common feature in LMXBs with high X-ray activity (i. e. steady systems and transients during outbursts). However, the NIII/CIII emission lines are excited on the inner hemisphere of the donor star by the Bowen mechanism/photoionization (\\citealt{MCT75}) and only provides a lower limit ($K_{em}$) to the true $K_{2}$-velocity of the donor. In \\cite{MCM05} we tackle this problem in a general approach by modeling the deviation between the reprocessed light-center and the center of mass of a Roche lobe filling star (the so-called 'K-correction') including screening effects by a flared accretion disc. In this paper we compute the K-correction for GX 339-4 which provides a more restrictive lower limit to the mass of the BH in this binary. In sections \\ref{companion} and \\ref{observations} we discuss the nature of the companion star and show that all the known observables can be explained by an scenario in which a 'stripped-giant' companion transfers mass onto a BH. We also apply this evolutionary model to XTE J1550-564 which shows observational properties very similar to GX 339-4. The stellar parameters of the companion in this system are in excellent agreement with a stripped-giant scenario. ", "conclusions": "We have applied the K-correction to the BH LMXB GX 339-4. By considering the limit case where the emission line is formed at the limb of the irradiated region of the companion we derive a solid lower limit to $M_{X}\\geq 6 M_{\\odot}$ including the error bars in $f(M)$. Here we have only assumed $M_2\\geq0.166 M_{\\odot}$, the lower limit allowed by the stripped giant model. We find that the stripped-giant evolutionary model explains the non-detection of the companion by SFC01 and the X-ray behaviour of the source. In particular, we propose $M_{2}\\ga0.3M_{\\odot}$, for which we predict $\\dot{M_{2}}$ large enough to explain the frequent X-ray outbursts displayed by this source. This limit results in $M_{X}\\ga7M_{\\odot}$. From the maximum mass solution we find $q\\leq0.125$. On the other hand, we have also shown that the stripped-giant model successfully explains the observable properties ($M_2$, $R_2$, $T_{eff}$ and $\\dot{M}_2$) of the akin LMXB XTE J1550-564.\\\\" }, "0801/0801.4448_arXiv.txt": { "abstract": "At the surface of the Sun acoustic waves appear to be affected by the presence of strong magnetic fields in active regions. We explore the possibility that the inclined magnetic field in sunspot penumbrae may convert primarily vertically propagating acoustic waves into elliptical motion. We use helioseismic holography to measure the modulus and phase of the correlation between incoming acoustic waves and the local surface motion within two sunspots. These correlations are modeled assuming the surface motion is elliptical, and we explore the properties of the elliptical motion on the magnetic field inclination. We also demonstrate that the phase shift of the outward propagating waves is opposite to the phase shift of the inward propagating waves in stronger, more vertical fields, but similar to the inward phase shifts in weaker, more inclined fields. ", "introduction": "\\label{introduction} Helioseismology uses the observed solar surface acoustic wavefield to construct images of the subsurface structure of the Sun. Of particular interest has been the three-dimensional (3D) modeling of time-distance \\cite{DJHP93} observations of travel-time shifts to deduce the subsurface structure of active regions \\cite{KDS00,ZK03}. Assuming the travel-time shifts are due to perturbations in the sound speed below the spot, a general consensus in the models has emerged consistent with sound-speed reductions (relative to surrounding quiet Sun) near the surface ($\\lesssim 4$ Mm) and enhancements up to 15 Mm below sunspots \\cite{KDS00,CBK06}. Using ring diagram analysis \\inlinecite{BAB04} also find a lower sound speed immediately below the surface and an increase in the sound speed below 7 Mm. A comparison between Fourier-Hankel analysis and time-distance results by \\inlinecite{B97} first prompted caution in the interpretation of acoustic oscillation signals within sunspots. The influences of strong surface magnetic fields have not been explicitly included in most helioseismic models of active regions. \\citeauthor{LB05ii} (2005a) and \\citeauthor{LB05i} (2005b) have shown that helioseismic phase shifts observed with helioseismic holography vanish below a depth of about 5 Mm, when a surface (``showerglass'') phase shift based on photospheric magnetic flux density, is removed from the data. Other evidence supports the possibility of strong near-surface contributions to the helioseismic phase (or travel-time) shifts. These include the possible contamination of surface perturbations into the 3D inversions (e.g. \\opencite{K06}; \\opencite{CR07}). It has also been shown that the reduction of $p$-mode amplitudes in magnetic regions can cause travel-time shifts \\cite{RBDTZ06}. The suppression of sources of wave excitation within sunspots can also produce measurable shifts \\cite{HCRB07}. \\inlinecite{SBCL05} and \\inlinecite{SBC07} have found that phase shifts obtained from seismic holography in sunspot penumbrae vary with the line-of-sight angle (from vertical) as projected into the plane containing the magnetic field and the vertical direction. A similar effect has also been noted by \\inlinecite{ZK06} with time distance measurements. \\inlinecite{SBC07} find that the effect is dependent upon the strength and/or inclination of the magnetic field. In the penumbra the magnetic field strength decreases as the magnetic field angle from vertical increases, hence the two properties of the magnetic field cannot be extricated. The phase variation with line-of-sight viewing angle is demonstrated most substantially at frequencies around 5 mHz with a strong, almost vertical magnetic field close to the umbra. \\inlinecite{SBC07} also find that the total variation across all lines-of-sight increases with temporal frequency, particularly in the stronger fields in the penumbrae. Mode conversion of the acoustic waves in the near surface has been explored as the physical cause of the observed absorption of acoustic waves by sunspots. A fast acoustic wave, propagating towards the surface from the interior, encounters the depth at which the Alfv\\'en speed is equal to the sound speed ($a\\approx c$) which is typically close to the surface in a sunspot. Under these conditions it is able to transmit to a slow acoustic mode and convert to a fast magnetic mode \\cite{C05}. \\inlinecite{CC03} explore the mode conversion in two-dimensions with a uniform \\emph{inclined} magnetic field relevant to sunspot penumbrae. The inclination of the magnetic field is found to have a significant dependence on the likelihood of conversion and fits extremely well with the analysis of \\inlinecite{B95} \\cite{CCB03}. Further work \\cite{C05,SC06} using ray theory has since established that it is the angle between the acoustic wave path and the magnetic field (the `attack angle') which is the crucial factor inducing conversion. With a wide attack angle at the $a\\approx c$ level there is maximum conversion from a fast acoustic to a fast magnetic mode. A fine attack angle encourages transmission to a slow acoustic mode, which is guided `up' the magnetic field lines to observational heights in the atmosphere. A consequence of mode conversion may be the observational signature of elliptical motion in regions of inclined magnetic field. The aim of this paper is to model the observations of two sunspots previously analyzed by \\inlinecite{SBC07} in order to determine the properties of velocity ellipses consistent with the data. These models are based on a least-squares-fit of the phase and modulus information of the local ingression control correlation. We explore the properties of these ellipses, observed with waves at different temporal frequencies, as functions of the magnetic field inclination angle. We also examine the variation with line-of-sight angle of the phase shifts in the outgoing waves, using the local egression control correlation, to assess the relation of the phase shift variations between incoming and outgoing waves. In the following sections we describe the data (Section \\ref{obs}), give an outline of the helioseismic holography technique used (Section 3), describe the results (Section 4) and discuss our results in the context of mode conversion (Section 5). ", "conclusions": "Mode Conversion predicts (among other things \\cite{C07}) that when the attack angle is small most of the observable energy will be in the slow acoustic mode, and when the attack angle is large most of the observable energy will be in the fast magnetic mode. In this case we will be seeing the line-of-sight effect of a combination of waves coming from all directions impinging on magnetic field with a particular orientation. The observations by MDI consist of line-of-sight Doppler signatures of the surface motion which are presumably caused mainly by pressure perturbations. This means that we would expect to be observing only the slow acoustic mode, however it is possible that we are observing a combination of the acoustic and magnetic modes. The dependence of the phase shift of the observed ingression correlations with azimuthal angle around sunspot penumbrae, as viewed from different observational vantages, shows that the incoming phase shifts must be (at least partly) photospheric in origin and are influenced by the presence of inclined magnetic fields \\cite{SBCL05}. Analysis of the variation of both the amplitude and phase of the surface velocities provides an opportunity to characterize the magnetically influenced acoustic signature as an ellipse with properties determined by the magnetic fields. We found that the ellipses are either nearly vertical (for weaker, more inclined fields) or generally directed \\emph{away} from the magnetic field direction (for stronger, more vertical fields). Largely consistent results for two active regions, AR9026 and AR9057, are found. Some properties of the surface ellipses, e.g. their inclinations, are different for the two sunspots. Some of this variation may be due to differences in the field properties. For example, the field in AR9057 is on average $\\sim 15\\%$ weaker than AR9026. Fits of the elliptical motion in Figures~\\ref{ellipse9026} and \\ref{ellipse9057} depend critically on the correlation modulus which is prone to systematic uncertainties. But the trend is that a stronger, less inclined magnetic field produces elliptical motion with smaller amplitude, eccentricity, and deviation angle, and a larger inclination from vertical. These trends exist for both spots and, largely, at all frequencies. The shorter semi-major axis at strong magnetic field strengths is consistent with previous knowledge of surface acoustic amplitude suppression in magnetic fields. It is curious to note, however, that at 4 mHz the amplitude is consistently smaller than even 3 mHz. These are the first results to estimate the behaviour of the surface velocity ellipse at the photosphere within sunspots. The results do not immediately suggest an observation of the slow wave as shown by \\inlinecite{C05} or \\inlinecite{SC06}, but are consistent with their expectations of the behaviour of slow waves at this height in the atmosphere. Mode conversion theory states the alignment is dependent upon $a^2/c^2$ which at the observational heights of $\\sim 200$km of the atmosphere may not be large enough to invoke alignment. Since the ray analysis is somewhat unrealistic we would expect to observe a combination of fast and slow waves at the surface, which will contribute to a clouded view of the surface velocities. \\inlinecite{RSWS07} are currently exploring the possibility that these apparent surface effects are due to the changes in the radiative transfer within active regions and the formation height of the observational Ni 678 nm line. This explanation requires an absorption mechanism, or else some other means of producing a difference between the amplitudes of upward and downward propagating waves. Thus mode conversion may still be important in this proposed mechanism. A test of the mechanism proposed by Rajaguru et al. (2007) would be to repeat the observations performed here in a magnetically insensitive line, where the proposed radiative transfer effects would not be present. In terms of mode conversion, it is suggested that the main effect occurs along the bright radial filaments of the interlocking comb structure as presented in the penumbral models of \\inlinecite{WTBT04}. However, observational helioseismic spatial resolution cannot currently resolve this. This is also the first time that the variation of the phase of the local egression correlation has been analysed in the penumbra. It is curious that the egression correlation shows a reverse dependence when the magnetic field is weak and highly inclined. This is evidence of a reverse ingression dependence on the line-of-sight, but further investigation is required to understand the behavior at high frequencies in the weaker, more inclined fields. \\vspace{2cm} This work was supported in part by the \\textit{European Helio- and Asteroseismology Network} (HELAS)." }, "0801/0801.2974_arXiv.txt": { "abstract": "General Relativistic (GR) Magnetohydrodynamic (MHD) simulations of black hole accretion find significant magnetic stresses near and inside the innermost stable circular orbit (ISCO), suggesting that such flows could radiate in a manner noticeably different from the prediction of the standard model, which assumes that there are no stresses in that region. We provide estimates of how phenomenologically interesting parameters like the ``radiation edge\", the innermost ring of the disc from which substantial thermal radiation escapes to infinity, may be altered by stresses near the ISCO. These estimates are based on data from a large number of three-dimensional GRMHD simulations combined with GR ray-tracing. For slowly spinning black holes ($a/M<0.9$), the radiation edge lies well inside where the standard model predicts, particularly when the system is viewed at high inclination. For more rapidly spinning black holes, the contrast is smaller. At fixed total luminosity, the characteristic temperature of the accretion flow increases between a factor of $1.2-2.4$ over that predicted by the standard model, whilst at fixed mass accretion rate, there is a corresponding enhancement of the accretion luminosity which may be anywhere from tens of percent to order unity. When all these considerations are combined, we find that, for fixed black hole mass, luminosity, and inclination angle, our uncertainty in the characteristic temperature of the radiation reaching distant observers due to uncertainty in dissipation profile (around a factor of 3) is {\\it greater} than the uncertainty due to a complete lack of knowledge of the black hole's spin (around a factor of 2) and furthermore that spin estimates based on the stress-free inner boundary condition provide an upper limit to $a/M$. ", "introduction": "Recent years have seen a rapid growth in the specificity and detail with which we attempt to describe accretion onto black holes. In the early days of the subject three decades ago, we were content with the simplest of models \\cite[the `standard' model, ][]{Novikov:1973}, hereafter NT, in which the accretion flow was assumed to be time-steady, axisymmetric, geometrically thin, and following circular Keplerian orbits at all radii outside the innermost stable circular orbit (the ISCO, located at radius $r_{ms}$). Relying on heuristic arguments framed in a purely hydrodynamic context \\citep{Page:1974}, it was further assumed that all stresses ceased inside the ISCO, so that the inner edge of the disc could be described as falling precisely at that radius. Dimensional analysis was the basis for any link between the inter-ring stresses essential to accretion and local physical conditions \\citep{Shakura:1973}. Today, there is tremendous effort to obtain and interpret direct observational diagnostics of the innermost parts of accretion flows onto black holes. Apparently relativistically-broadened Fe K$\\alpha$ profiles can be discerned in numerous examples \\citep{Reynolds:2003}. Extensive efforts are made to fit detailed disc spectral models to observed continuum spectra \\citep{Gierlinski:2004,Davis:2005,Shafee:2006a,Shafee:2006b,Hui:2007}. These spectral fits can be used to infer the mass of the central black hole; both methods may be used to constrain the spin of the black hole \\citep{Makishima:2000,Miller:2004,Miniutti:2004,Brenneman:2006,Shafee:2006a, Shafee:2006b}. Some hope to use relativistic fluid dynamics to define normal modes of disc oscillation that could then also be used to constrain the central black hole's spin \\citep{Wagoner:2001,Rezzolla:2003,KatoS:2004}. At the same time, there has also been much progress on the theoretical side. A strong consensus now supports the idea that accretion stresses are the result of correlated MHD turbulence, driven by a pervasive magneto-rotational instability \\citep{Balbus:1998}. Numerical simulations developing this idea have given us a detailed and quantitative description of the vertical profiles of pressure and density inside discs \\citep{Miller:2000,Hirose:2005,Hirose:2007}, as well as provided us with detailed pictures of how their properties vary smoothly across the ISCO \\citep{Krolik:2005}. These simulations have vindicated the prescient remark made by \\citet{Thorne:1974}: \\noindent ``In the words of my referee, James M. Bardeen (which echo verbal warnings that I have received from Ya. B. Zel'dovich and V.F. Schwartzman), `It seems quite possible that magnetic stresses could cause large deviations from circular orbits in the very inner part of the accretion disk and change the energy-angular-momentum balance of the accreting matter by an amount of order unity'.\" Indeed, when the black hole rotates, significant magnetic stresses can be found throughout the accretion flow, all the way to the edge of the event horizon \\citep{Krolik:2005}. Work on specific dynamical pictures of accretion has stimulated a reexamination of the simple picture that discs have sharp edges, within which little of interest happens. Although it is true that there are qualitative changes across the ISCO, they do not happen discontinuously. As argued by \\cite{Krolik:2002}, what one means by ``edge\" depends on the question asked. For example, the ``reflection edge\", the edge outside of which most of the observed Fe K$\\alpha$ and Compton reflection photons are created, is likely to lie near, but possibly either inside or outside, the ISCO \\citep{Reynolds:1997,Krolik:2002,Brenneman:2006,Reynolds:2008}---its exact position depends on the density and optical depth of the gas in that region, and on the intensity of ionizing radiation striking it. Similarly, the ``radiation edge\" (the edge inside of which little of the total luminosity emitted by the flow escapes to infinity) should be near the ISCO, but is not necessarily identical to it. Its position depends both on the profile of dissipation and on the ability of photons to escape to infinity, and to do so without excessive loss of energy to gravitational redshift. In this paper we examine the influence of magnetic stresses at and inside the ISCO on the apparent size, luminosity and characteristic temperature of black hole accretion discs. This effort is important to black hole phenomenology because so many observational diagnostics depend upon these three parameters. A prime example is provided by attempts to use spectral fits to constrain black hole mass and spin. This program rests upon the idea that the observed luminosity, $L$, is essentially thermal, so that it may be characterized by an effective radiating area and a characteristic temperature: $L = A(a/M,\\theta) T^{4}_{char}$. Because both $T_{char}$ and $L$ can be measured, $A$ can be inferred through this relation. With a theoretically-supported connection between $A$ and $a/M$ (and some other constraint on $\\theta$), the inference of $A$ leads to an inference of $a/M$. Current efforts connect $T_{char}$ to $M$, $\\dot M$, and $a/M$ using the \\cite{Novikov:1973} model for the radial dissipation profile. This model depends in an essential way on the assumption that all stresses cease at and within the ISCO, whose location (at Boyer-Lindquist radial coordinate $r/M=r_{ms}$) is a function only of the black hole mass and spin. This temperature is further adjusted by an appropriate correction for gravitational and Doppler energy shifts and a ``color temperature correction'' due to opacity effects \\cite[see e.g.][]{Done:2008}. The apparent size of the disc, $A$, is likewise found from the dissipation profile of the Novikov-Thorne model. As first noted by \\cite{Page:1974}, significant magnetic forces undercut the rationale for the zero-stress boundary condition; the simulation data we discuss here shows quantitatively how these magnetic forces alter the connections between both $A$ and $T_{char}$ and the black hole spin parameter. Although more work is needed to explore fully these new effects, in this paper we begin the discussion of how they can influence these inferences. The simulations reported here employ full general relativity and three-dimensional MHD, so that whilst their treatment of angular momentum flow and inflow dynamics is quite accurate, they do not directly track dissipation. In an accretion disc, energy is extracted from orbital motion and transformed into kinetic and magnetic energy on the largest scales of the turbulence. Subsequently, this energy cascades down to a dissipation scale (either viscous or resistive) where it is finally thermalized. Current simulations can describe well the first stages of this process, but can mimic only indirectly the last step: grid-level effects intervene at lengthscales far larger than the physical scale of dissipation. In fact, the simulation code whose data we will use solves only the internal energy equation, and makes no attempt to follow dissipative energy losses except those associated with shocks. We proceed by instead making a plausible {\\it ansatz} for heating within the disc that can be determined \\textit{a posteriori } from simulation data. As we shall see, our results depend primarily on the qualitative fact that dissipation continues smoothly across the ISCO, and on the nature of photon trajectories deep in a relativistic potential; for this reason, we believe that a non-rigorous, but physically-motivated, {\\it ansatz} will not be misleading. We opt for the simplest choice: a connection between the heating rate and the stress that follows from the standard model for energy conservation in an accretion disc \\citep{Page:1974,Balbus:1994,Hubeny:1998}. We can also investigate the importance of enhanced stress in a semi-analytic fashion that bypasses most of the limitations of the current simulations. We employ the model formulated by \\cite{Agol:2000} (hereafter the AK model) that allows for non-zero stress at the ISCO but otherwise computes the dissipation profile using the same approach as the standard model. The AK model cannot be extrapolated into the plunging region, and has in common with the NT model a fixed disc size. Its unique feature is the enhanced total dissipation due to the nonzero stress at the ISCO; simulation data provides the single parameter needed to calibrate the model. The AK model, therefore, provides an important link between the standard model and the full simulation results and allows us to gauge the appropriateness of the radiation {\\it ansatz} employed for the latter. \\begin{table*} \\caption{Simulation Parameters} \\label{sims} \\begin{tabular}{@{}lccccccc} \\hline Name & $a/M$ & $r_{+}/M$ & $r_{\\rm in}/M$ & $r_{\\rm ms}/M$ & Field & $T_{avg}\\times10^{3}M$ & Originally Presented in \\\\ \\hline KD0b & 0.0 & 2.00 & 2.104 & 6.00 & Dipole & 6--8 & \\cite{De-Villiers:2003b} \\\\ KD0c & 0.0 & 2.00 & 2.104 & 6.00 & Dipole & 8--10 & \\cite{Hawley:2006} \\\\ QD0d & 0.0 & 2.00 & 2.104 & 6.00 & Quadrupole & 8--10 & This work \\\\ KDIa & 0.5 & 1.86 & 1.904 & 4.23 & Dipole & 6--8 & \\cite{De-Villiers:2003b} \\\\ KDIb & 0.5 & 1.86 & 1.904 & 4.23 & Dipole & 8--10 & \\cite{Hawley:2006} \\\\ KDPd & 0.9 & 1.44 & 1.503 & 2.32 & Dipole & 6--10 & \\cite{De-Villiers:2003b} \\\\ KDPg & 0.9 & 1.44 & 1.503 & 2.32 & Dipole & 8--10 & \\cite{Hawley:2006} \\\\ QDPa & 0.9 & 1.44 & 1.503 & 2.32 & Quadrupole & 8--10 & \\cite{Beckwith:2008a} \\\\ TDPa & 0.9 & 1.44 & 1.503 & 2.32 & Toroidal & 20--22 & \\cite{Beckwith:2008a} \\\\ KDG & 0.93 & 1.37 & 1.458 & 2.10 & Dipole & 8--10 & \\cite{Hawley:2006} \\\\ KDH & 0.95 & 1.31 & 1.403 & 1.94 & Dipole & 8--10 & \\cite{Hawley:2006} \\\\ KDJd & 0.99 & 1.14 & 1.203 & 1.45 & Dipole & 8--10 & This work \\\\ KDEa & 0.998 & 1.084 & 1.175 & 1.235 & Dipole & 6--8 & \\cite{De-Villiers:2003b} \\\\ KDEb & 0.998 & 1.084 & 1.175 & 1.235 & Dipole & 8--10 & This work \\\\ QDEb & 0.998 & 1.084 & 1.175 & 1.235 & Quadrupole & 8--10 & This work \\\\ \\hline \\end{tabular} \\medskip Here $a/M$ is the spin parameter of the black hole, $r_{+}$ is the horizon radius, $r_{\\rm in}$ is the innermost radius in the computational grid, field is the initial field topology in the torus and $T_{avg}\\times10^{3}$ is the time-interval over which simulation data was averaged. For reference, we note where individual simulations were originally presented. \\end{table*} To relate dissipation rates to radiation received at infinity, we must perform one additional calculation using three further assumptions: that the dissipated heat is efficiently converted to photons, that these photons emerge from the accretion flow very near where they are created, and that they are radiated isotropically in the fluid frame. The first two assumptions are equivalent to requiring the timescales for dissipation, radiation, and photon diffusion to be shorter than the inflow timescale for all fluid elements. The third, while not strictly justified, is the simplest guess we can make. Given those assumptions, we use a general relativistic ray-tracing code to relate the luminosity radiated by each fluid element to the luminosity received by observers located at different polar angles far from the black hole. A further result of this calculation is a new estimate of the radiative efficiency of accretion. The traditional calculation of this quantity follows directly from a primary assumption that there are no forces inside the ISCO and two additional assumptions, that the radiation is prompt and all of it reaches infinity. We improve upon this traditional estimate in two ways: we allow for dissipation associated with the stresses we measure at and inside the ISCO; and we calculate the radiated energy (even within the NT model) that actually reaches distant observers. However, we do not regard our result as sufficiently final or complete to give it much weight, as it is likely to be more model- and parameter-dependent than our placement of the radiation edge. One reason for downplaying this result is that we find it necessary to omit any estimate of dissipation outside the main part of the accretion flow. We have less confidence that our dissipation prescription is appropriate in the jet, or even in the disc corona, than we do when it is applied to the main disc body and plunging region. Moreover, radiation from these lower-density regions is less likely to be thermalised and contribute to the radiation usually identified with the disc continuum. The rest of this paper is structured as follows. In \\S\\ref{numerics} we briefly review the numerical scheme employed to solve the equations of GRMHD in the simulations whose data we use, give an overview of the parameters of the simulations included in this work, and describe our general relativistic ray-tracing code. In \\S\\ref{dissmodels}, we contrast the dissipation rate distributions predicted by the standard model, its Agol \\& Krolik extension, and our simulation-based {\\it ansatz}. In \\S\\ref{rad} we compute the luminosity at infinity predicted by each of these dissipation profiles and discuss their consequences for the location of the radiation edge and the characteristic temperature of the accretion flow. Finally, in \\S\\ref{summ}, we draw specific conclusions from our results and describe their implications for black hole phenomenology. ", "conclusions": "\\label{summ} That magnetic forces might cause substantial stress at the ISCO was foreseen very shortly after the invention of the standard model \\cite[][]{Page:1974}. This possibility now appears to be an immediate corollary of the well-established result that MHD stresses account for most of the angular momentum transport in the bodies of accretion discs. Indeed, such stresses were seen in the first generation of three dimensional GRMHD disc simulations. The goal of this paper has been to begin the linkage of these numerical MHD simulations to the observable properties of accreting black holes even before the simulations are fully equipped to make predictions about how these systems radiate. To do so, we have followed a path of cautious extrapolation from older methods. We first used simulation data to fix the single parameter of a model (called AK here) that changes the previous standard (the Novikov-Thorne model) only by admitting the possibility of a non-zero stress at the ISCO. Because the AK model is defined in a way that prevents its extrapolation within the ISCO, we used the formalism underlying both it and the NT model (i.e., vertically-integrating and azimuthally- and time-averaging the equation of momentum-energy conservation under the assumption that the four-velocity and the stress tensor are orthogonal) to create an expression for the dissipation (called $Q_{\\mathrm{MW}}$ here) valid both inside and outside the ISCO. \\begin{figure} \\begin{center} \\includegraphics[width=0.48\\textwidth]{saavg_tchar} \\end{center} \\caption[]{Plot showing solid angle averaged $\\{ T_{char} \\}$ . Symbols are as in Figure \\ref{dissedge}} \\label{saavgtchar} \\end{figure} Happily, in the body of the accretion disc well outside the ISCO, both the AK and the MW method agree fairly well with NT more or less independent of black hole spin and magnetic field topology, although the irregularity in the curves of Figure~\\ref{spindiss} reminds one that the simulations are dynamic and time-varying, and that the 26 samples of simulation data we used define only somewhat imperfectly the long-term time-average. In addition, with the exception of the extreme high-spin example, where the AK and MW methods depart from NT just outside the ISCO, they do so together. This is noteworthy because, although they are based on closely-related formalisms, they are not identical: perhaps their most significant contrast is that the AK method assumes that $u_r = 0$, while the MW method does not. Lastly, inside the ISCO, where only the MW method is defined, it follows a smooth extrapolation from larger radius. When the black hole spins slowly ($a/M \\lesssim 0.9$), $Q_{\\mathrm MW}$ extends with hardly any change in logarithmic derivative with respect to radius. For higher spin, the extension gradually steepens toward smaller radius, but the next step in our formalism shows that this makes little difference to observed radiation: relativistic ray-tracing shows that the volume deep inside the plunging region, particularly at high spin, contributes little energy to the luminosity reaching observers at infinity. Thus, we are relatively confident that, despite the uncertainties involved, our estimate of the location of the radiation edge is comparatively insensitive to the exact relation between the flow's detailed dynamical properties and the dissipation rate. \\begin{figure} \\begin{center} \\includegraphics[width=0.48\\textwidth]{tcshape} \\end{center} \\caption[]{Plot showing the region of the $(T_{char},a/M)$-plane consistent with a given value of $\\theta_{o}$ for a $10 M_{\\odot}$ black hole accreting at $0.1L_{edd}$. Different colors correspond to different $\\theta_{o}$ (see key in bottom left corner of plot), dot-dash lines show the $T_{char}$ curve predicted by $Q_{NT}$ at a given $\\theta_{o}$, and the region enclosed by the dot-dash and dashed lines of a given color show the region of the $(T_{char},a/M)$-plane consistent with that value of $\\theta_{o}$.} \\label{tcshape} \\end{figure} The dependence of the radiation edge on spin may be summarised succinctly: At the highest spin, there is relatively little difference between the different methods of estimating its position because the ISCO is so close to the horizon that the great majority of photons released in the plunging region never reach infinity, or if they do, are severely redshifted. It moves from 2--3 times the radius of the ISCO when the disc is viewed face-on to almost exactly at the ISCO when the disc is viewed nearly edge-on. This inward movement of the radiation edge with increasing inclination angle is quite model-independent, as it stems from relativistic photon propagation effects: when photons from the plunging region do reach infinity with substantial energy, it is because they are emitted in the direction of the orbital motion. Probability of escape is then enhanced by a combination of special relativistic beaming and gravitational lensing; energy at infinity is enhanced by special relativistic Doppler boosting. As the black hole spin decreases, the diminishing depth of the potential immediately inside the ISCO makes it progressively easier for photons to escape from that region and reduces the gravitational redshift they suffer when they do. The result is that the radiation edge moves farther inside the ISCO as {\\it either} the spin diminishes (at fixed viewing angle) or the inclination angle moves toward the equatorial plane (at fixed spin). At its most extreme, the case of $a/M = 0$ and $\\theta_o = 85^{\\circ}$, $r_{re}$ can be $\\simeq 0.5 r_{ms}$. Figure~\\ref{ofrms} gives additional cause to believe that these results are comparatively insensitive to dissipation model. Although the radiation edge can move well inside the ISCO at low spin and high inclination, most of the light received by distant observers is generally emitted in the region near and outside the ISCO. Only at the highest inclinations ($\\theta_o \\gtrsim 75^{\\circ}$) and lowest spins ($a/M \\lesssim 0.5$) does the contribution of the plunging region to the luminosity approach $50\\%$. Thus, most of the light seen at infinity likely comes from a region where the predictions of the AK and the MW models differ little. Nonetheless, because it is also true that most of the light is emitted within a radius at most a few times the ISCO (except for the highest spin viewed more or less face-on), the contrast in total luminosity between the AK and MW models on the one hand, and the NT on the other, are order unity for all spins $\\leq 0.9$. For higher spins, the effect may be smaller, but the uncertainties are also greater. These conclusions have immediate implications for a number of phenomenological issues. Firstly, as suggested by \\cite{Falcke:2000}, it may be possible to image the nearest supermassive black hole, the one in Sgr A*. Because its accretion flow, unlike those of intrinsically brighter systems, could well be radiatively inefficient, a simulation scheme that conserves total energy is more appropriate to analysing its emission. \\cite{Noble:2007}, using such a code (albeit an axisymmetric version), have produced predicted images that illustrate several of the effects emphasised here, although in their work so far they have not reported quantitative descriptions of characteristic emission radii. Secondly, the enhanced total radiative efficiency due to dissipation in the marginally stable region may affect estimates of population-mean spin parameters \\cite[e.g., as for AGN by][] {Elvis:2002,Yu:2002}. Because the efficiency rises with increasing prograde spin in the NT model, the spin inferred by this method may overestimate the actual spin of accreting black holes if this enhancement is ignored. The additional luminosity from enhanced stress in the innermost part of the accretion flow could significantly alter the emergent spectrum. Employing a simple thermal model, we have found that the characteristic temperature of the flow increases by a factor of 1.2 --1.4 over that predicted by the NT model. As a consequence, the thermal peak of the disk spectrum (at $\\sim 1$~keV in Galactic black holes, $\\sim 10$~eV in AGN) may be pushed to somewhat higher energies. Several caveats must be mentioned, however, in regard to this prediction. First, this number supposes an emergent spectrum that is Planckian, but most estimates of the disc atmosphere's structure suggest that it is scattering-dominated, so that the color temperature of the spectrum is shifted upward from the effective temperature. The magnitude of this shift depends on details of the disc's vertical structure that are not as yet well known \\cite[see e.g.][]{Davis:2005}. Furthermore, \\cite{Blaes:2006} \\citep[using the vertically stratified shearing box simulations of][]{Hirose:2005} show that magnetic pressure support changes the vertical structure of the disk resulting in a noticeable hardening of the emergent disk spectrum compared to the standard Novikov-Thorne picture due to non-LTE effects. Second, it is possible that some of the enhanced dissipation will occur where the density and optical depth are too low to accomplish thermalisation. Strengthening of the ``coronal\", i.e., hard X-ray, emission, rather than hardening the thermal disc spectrum would then be the likely consequence. Third, our treatment ignores those photons emitted deep in the potential that neither escape directly to infinity nor are captured by the black hole, but instead strike the disc. As shown by \\cite{Agol:2000}, this ``returning radiation\" can be a substantial fraction of all photons emitted when $r \\lesssim 5r_g$. Depending on their spectrum and the structure of the disc atmosphere where they strike, these photons may be either reflected (with Doppler shifts) or absorbed and their energy reradiated at a different (in general, lower) temperature. Quantitatively evaluating all three of these effects is well beyond the scope of this paper, but can be done in future work. There are also implications for attempts to determine black hole spin from spectral fitting. In all three models, the characteristic radius of emission is always {\\it near} the ISCO, but does not, in general, coincide with it. Generally speaking, this characteristic radius is largest for the NT model, smaller (but still outside the ISCO) for the AK model, and smaller still, possibly moving into the plunging region inside the ISCO, for the MW model. Because the ISCO moves to smaller radial coordinate as $a/M$ increases, these characteristic radii always become smaller for faster spin. However, the fractional amount by which the characteristic emission radius moves inward in the MW model is greatest for the lowest spins, so that in the end, the MW model predicts a relatively slow variation of radiation edge with black hole spin. The AK model, like the NT model, does not radiate from inside the ISCO, but the additional stress at and just outside the ISCO in this model (relative to the NT prediction) produces a systematic increase in the characteristic temperature. The magnitude of this shift in characteristic temperature rises, of course, with increasing additional stress. When there is emission from the plunging region, as in the MW model, the characteristic temperature can rise still higher, but the highly relativistic motions there make observed properties more strongly dependent on inclination angle. In addition, a larger fraction of the emitted photons can be captured by the black hole. When all these considerations are combined, we find that, for fixed black hole mass, luminosity, and inclination angle, the uncertainty in the characteristic temperature of the radiation reaching distant observers due to uncertainty in the dissipation profile is {\\it greater} than the that due to a complete lack of knowledge of the black hole's spin. Clearly, our incomplete understanding of accretion disc physics (here specifically the magnitude of the stress at and inside the ISCO) makes it difficult to determine a black hole spin based on continuum model-fitting. The best one can say is that estimates based on the traditional Novikov-Thorne model can be expected to yield the most rapid spin possible, but the actual spin may be significantly slower. Our results demonstrate the potential importance of nonzero stresses at and inside the ISCO. But how representative are the specific values obtained in these simulated discs? There are two considerations: those arising from purely numerical effects, and those limitations arising from the assumptions and parameters of the model used. First, the results of numerical simulations can be influenced by finite resolution and the limitations of the numerical technique. All of the simulations presented in this work were performed at a resolution of $192\\times192\\times64$ $(r,\\theta,\\phi)$ grid zones using ideal MHD and an internal energy equation. The equation of state and the numerical energy dissipation are unlikely to have a direct effect on our conclusions as $Q_{MW}$ is derived directly from the \\emph{physical} Maxwell stresses within the disc, rather than by measuring some \\emph{numerical} dissipation rate. Low resolution usually causes the Maxwell stress to be {\\it undervalued}; if so, the implications of this paper would be strengthed by improved resolution. Until available computer power makes better-resolved three-dimensional simulations possible, the best way we have to test the effects of finite resolution is to compute axisymmetric simulations with higher resolution. A variety of such simulations were presented in \\cite{Beckwith:2008a} with resolutions up to $1024^2$. We observe that greater resolution reduces the rate of numerical reconnection and improves the ability of the simulation to maintain certain field configurations and small-scale field structures. Overall the amplitude of the turbulent Maxwell stresses in the disc remained largely unchanged as resolution was increased. We have also calculated ${\\cal W}^{(r)}_{(\\phi)}$ and $Q_{MW}$, and find no significant qualitative differences from the results presented in this work. \\newpage Beyond the purely numerical issues, the value of the Maxwell stress at the ISCO may depend on a number of disc properties. In the ensemble of simulations presented here, the stress levels are determined in part by the initial field topology (dipole versus quadrupolar, poloidal versus toroidal, and the presence or absence of a net vertical field). Indeed, local shearing-box simulations suggest that the saturated field strength can increase substantially when large-scale vertical field threads the disc \\cite[][]{Balbus:1998}. It is also possible that the saturation stress depends on disc thickness. To quantify disc thickness, we define the scale-height $H$ as the proper height above the plane at which the time- and azimuthally-averaged density falls by $1/e$ from its similarly averaged value on the equatorial plane (see \\S\\ref{dissmodels}): $H= \\int^{\\theta_h}_{\\pi/2} \\sqrt{g_{\\theta \\theta} (r=3M,\\theta)} d\\theta$. We similarly define $R$ as the proper (as opposed to coordinate) radial distance from the horizon to $3M$ plus the coordinate distance from the origin to the horizon: $R = r_{in} + \\int^{r=3M}_{r_{in}} \\sqrt{g_{rr} (r,\\theta=\\pi/2)} dr$, with $r_{in}$ as given in Table \\ref{sims}). We then define the disc thickness as the ratio $H/R$. Even though $r=3M$ lies well within the plunging region for the lower spin cases, we find that there is a slow enough radial variation in this quantity to make $H/R$ a reasonably well-defined parameter. Measured in this way, our discs are modestly thick, with a characteristic aspect ratio $H/R = 0.06$--0.2 at $r=3M$. Most of the range in $H/R$ results from the fact that the maximum pressure in the initial condition for these simulations varies somewhat between different $a/M$. It is difficult to say, however, just what sort of dependence there may be on disc thickness. There are some arguments suggesting that magnetic effects may increase with increasing $H/R$. For example, local shearing-box simulations find that the Maxwell stress is proportional to magnetic pressure \\cite[][]{Balbus:1998,Sano:2004}, but there have not yet been any systematic studies of what regulates the magnetic pressure in global, vertically-stratified discs. \\cite{Afshordi:2003} have argued that inner disc stresses and dissipation may depend on disc thickness as well as on accretion rate, an argument reiterated by \\cite{Shafee:2007}, but their arguments are framed in an essentially hydrodynamic context, and therefore eliminate any possibility of predicting magnetic stresses. They are also non-relativistic, and therefore eliminate any effects due to frame-dragging. There are also arguments that any dependence on $H/R$ may be weak. In the simplest analytic or quasi-analytic MHD models, magnetic torques at the ISCO remain significant even in the limit of a zero pressure disc \\citep{Krolik:1999a,Gammie:1999}. As discussed in \\cite{Krolik:2005}, processes analogous to the Blandford-Znajek mechanism can readily transport energy and angular momentum from rotating black holes to the accretion flow, and there is no particular reason to think that these processes should be tightly connected to the pressure in the disc. In the end, only direct simulations with the resolution to describe thin discs adequately will answer this question in a satisfactory way. The results of the present investigation provide yet one more reason why it will be important to do so. The final question that we address in this work is how the results presented here relate to current state of the art measurements of black hole spin via spectral fitting of the disk continuum. The 6 systems with the best data \\cite[see e.g.][]{Davis:2006,Shafee:2006a,Middleton:2006,Liu:2008} all have spins in the range $a/M\\sim0.1-0.8$ based on disk models that assume the stress-free inner boundary condition. From the perspective of the disk stress models these spins are upper limits. This might indicate that the hole is counter-rotating, but also opens the possibility that spin determinations might themselves constrain the physics near the ISCO. Firstly, the stress levels at the ISCO could be near the value assumed by the stress-free inner boundary condition; secondly, the classical relationship between stress and dissipation might not hold for enhanced magnetic stresses ear the ISCO; thirdly, the density levels at and inside the ISCO could be insufficient to thermalize the dissipated heat; fourthly, the time-scales for thermalization and radiation of the dissipated heat could be longer than the inflow time-scale. Another uncertainty which we have not examined is that the plane of the disk and the equatorial plane of the hole could be misaligned and so the disk is subject to the Bardeen-Peterson effect \\cite[see e.g.][]{Fragile:2007}. Understanding the roles played by the these effects will be crucial in providing robust estimates of black hole spins via spectral fitting of the disk continuum." }, "0801/0801.0189_arXiv.txt": { "abstract": "We study baryon number violating nucleon decays induced by unparticle interactions with the standard model particles. We find that the lowest dimension operators which cause nucleon decays can arise at dimension $6 + (d_s-3/2)$ with the unparticles being a spinor of dimension $d_s=d_\\U +1/2$. For scalar and vector unparticles of dimension $d_\\U$, the lowest order operatoers arise at $6+d_\\U$ and $7+d_\\U$ dimensions, respectively. Comparing the spinor unparticle induced $n \\to O^s_\\U$ and experimental bound on invisible decay of a neutron from KamLAND, we find that the scale for unparticle physics is required to be larger than $10^{10}$ GeV for $d_\\U < 2$ if the couplings are set to be of order one. For scalar and vector unparticles, the dominant baryon number violating decay modes are $n\\to \\bar \\nu + O_\\U (O^\\mu_\\U)$ and $p \\to e^+ + O_\\U (O^\\mu_\\U)$. The same experimental bound puts the scales for scalar and vector unparticle to be larger than $10^{7}$ and $10^{5}$ GeV for $d_\\U <2$ with couplings set to be of order one. Data on $p \\to e^+ \\mbox{invisible}$ puts similar constraints on unparticle interactions. ", "introduction": " ", "conclusions": "" }, "0801/0801.0006_arXiv.txt": { "abstract": "{ Small universe models predicted a cutoff in large-scale power in the cosmic microwave background (CMB). This was detected by the Wilkinson Microwave Anisotropy Probe (WMAP). Several studies have since proposed that the preferred model of the comoving spatial 3-hypersurface of the Universe may be a Poincar\\'e dodecahedral space (PDS) rather than a simply connected, flat space. Both models assume an FLRW metric and are close to flat with about 30\\% matter density. } { We study two predictions of the PDS model. (i) For the correct astronomical positioning of the fundamental domain, the spatial two-point cross-correlation function $\\ximc$ of temperature fluctuations in the covering space (where the two points in any pair are on different copies of the surface of last scattering (SLS)) should have a similar order of magnitude to the auto-correlation function $\\xisc$ on a single copy of the SLS. (ii) Consider a ``generalised'' PDS model for an {{\\em arbitrary}} ``twist'' phase $\\phi \\in \\left[0,2\\pi\\right]$. The optimal orientation and identified circle radius for a generalised PDS model found by maximising $\\ximc$ relative to $\\xisc$ in the WMAP maps should yield one of the two twist angles $ \\pm 36\\ddeg$. } { Comparison of $\\ximc$ to $\\xisc$ extends the identified circles method, using a much larger number of data points. We optimise the ratio of these functions at scales $\\ltapprox 4.0${\\hGpc} using a Markov chain Monte Carlo (MCMC) method over orientation ($l$, $b$, $\\theta$), circle size $\\alpha$, and twist $\\phi$. } { Both predictions were satisfied: (i) An optimal generalised PDS solution was found for two different foreground-reduced versions of the WMAP 3-year all-sky map, both with and without the kp2 galactic contamination mask. This solution yields a strong cross-correlation between points which would be distant and only weakly correlated according to the simply connected hypothesis. The face centres are $\\{(l,b)\\}_{i=1,6}\\approx \\{ (184\\ddeg, 62\\ddeg), (305\\ddeg, 44\\ddeg), (46\\ddeg, 49\\ddeg), (117\\ddeg, 20\\ddeg), (176\\ddeg, -4\\ddeg), (240\\ddeg, 13\\ddeg) \\}$ (and their antipodes) to within $\\approx 2\\ddeg$ ; (ii) This solution has twist $\\phi= (+39 \\pm 2.5)\\ddeg$, in agreement with the PDS model. The chance of this occurring in the simply connected model, assuming a uniform distribution $\\phi \\in [0,2\\pi]$, is about {\\probifnotPDS}. } { The PDS model now satisfies several different observational constraints. } ", "introduction": "\\label{s-intro} The past decade and a half have shown considerable growth in attempts to determine the global shape, i.e. not only the curvature, but also the topology, of the spatial comoving section of the Universe, i.e. of the 3-manifold to which a 3-hypersurface corresponds, informally known as ``space''. As noted by several authors, in particular \\nocite{Star93}{Starobinsky} (1993) and \\nocite{Stevens93}{Stevens} {et~al.} (1993), a space that is ``small'' compared to the surface of last scattering (SLS) cannot contain eigenmodes, which are used for expressing density perturbations, which are themselves larger than the space itself. This should lead to a cutoff of power in statistics representing these fluctuations, above which power should drop to zero. This prediction was made after COBE data were available, but {\\em before} the WMAP satellite was launched. For practical, observational reasons, spherical harmonic analyses of temperature fluctuations on the 2-sphere are frequently made. However, a physically more natural statistic to use is one in three-dimensional space, e.g. the two-point auto-correlation function. \\fauto The predicted cutoff in large-scale power appears to have been confirmed by the first-year observations of the Wilkinson Microwave Anisotropy Probe (WMAP) experiment. With this data, \\nocite{WMAPSpergel}{Spergel} {et~al.} (2003) published a figure approximately equivalent to such a function, i.e. the black ``WMAP data'' curve of Fig.~16 in their paper. Their figure shows the auto-correlation calculated as a function of angular separation, shown against projected spatial separation for a (first year) template-cleaned V map with the kp0 galactic contamination mask. The authors note the surprisingly flat correlation on large scales and suggest a multiply connected universe model to match this function. In Fig.~\\ref{f-auto}, we calculated the auto-correlation function directly as a function of three-dimensional spatial separation, not of angular separation, using the 3-year integrated linear combination (ILC) map with the kp2 cut. \\posteditorchanges{This figure can be approximately compared to the black ``WMAP data'' curve of Fig.~16 of \\protect\\nocite{WMAPSpergel}{Spergel} {et~al.} (2003), except that the present figure shows the auto-correlation calculated as a function of three-dimensional spatial separation, not of angular separation; \\langed{``however'' would be wrong here; 3-d space vs angle already does imply non-linearity, but this needs to be emphasised.} the relation between the two is {\\em not} linear. Also, our figure uses the 3-year ILC map with the kp2 cut, not the 1-year template-cleaned V map with the kp0 cut.} In spatial comoving units, we confirm that the auto-correlation is very close to flat for separations larger than $\\approx 10 {\\hGpc}$. The relation between angular and spatial scales is, of course, not linear. Equation~(\\ref{e-alpha-tri}) below (for either a spherical covering space or for a flat covering space using the limit $R_C \\gg \\rSLS \\ge d/2$) can be used to calculate this. If the size of the Universe\\footnote{See e.g. Fig~10 of \\protect\\nocite{LR99}{Luminet} \\& {Roukema} (1999) for a schematic diagram of various definitions of the ``size'' of the fundamental domain, including the injectivity radius, the in-radius, and the out-radius.} is about 10{\\hGpc}, as this figure seems to indicate, then which of the various 3-manifolds correctly describes comoving space? Motivated by indications that the Universe may have positive curvature, and using eigenmode-based simulations to study the spherical harmonic ($C_l$) spectrum of the WMAP data, \\nocite{LumNat03}{Luminet} {et~al.} \\nocite{LumNat03,Caillerie07}({Luminet} {et~al.} 2003; {Caillerie} {et~al.} 2007) argue that the Poincar\\'e dodecahedral space (PDS) is favoured by the WMAP data over an infinite, simply connected flat space. \\nocite{Caillerie07}{Caillerie} {et~al.} (2007) state that, by requiring maximal repression of the quadrupole signal, an optimal total density of $\\Omtot = 1.018$ is favoured (for a non-relativistic matter density $\\Omm \\equiv 0.27$ and Hubble constant $H_0 = 70$\\kms/Mpc) for the PDS model. Several other authors \\nocite{Aurich2005a,Aurich2005b,Gundermann2005}({Aurich} {et~al.} 2005a, 2005b; {Gundermann} 2005) have also compared simulations for PDS models to the observed first-year and three-year Wilkinson Microwave Anisotropy Probe (WMAP) maps of the cosmic microwave background (CMB). In all these studies, both the infinite flat models and the PDS models are used in the context of a standard, hot big bang model, i.e. where the Universe has a Friedmann-Lema\\^{\\i}tre-Robertson-Walker (FLRW) metric, perturbed by fluctuations that collapse gravitationally to form structures such as filaments and clusters of galaxies, and where values of the metric parameters consistent with the consensus obtained during the last decade of observations are adopted: the Universe is close to flat on length scales up to the SLS and has about 30\\% non-relativistic matter density. In other work, \\nocite{RLCMB04}{Roukema} {et~al.} (2004) used the identified circles principle \\nocite{Corn96,Corn98b}({Cornish} {et~al.} 1996, 1998) to find a specific optimal orientation of the PDS model based on the WMAP first-year ILC map, and published a tentative set of coordinates. \\nocite{KeyCSS06}{Key} {et~al.} (2007) confirmed the presence of a signal at the celestial coordinates, circle radius, and $-36\\ddeg$ twist published in \\nocite{RLCMB04}{Roukema} {et~al.} (2004), but argue that it should be considered a false positive. Using software independent of that used in \\langed{We want to avoid ambiguity between independence of the article Roukema et al. 2004 (true) and independence of the people Roukema et al. (false).} \\nocite{RLCMB04}{Roukema} {et~al.} (2004), updating to the 3-year WMAP data, and using Gaussian simulations, \\nocite{LewRouk2008}{Lew} \\& {Roukema} (2008) find similar conclusions. They find that a local maximum in the statistic \\langed{``'Statistics' is usually plural, unless this is a very specific mathematical or physical sort. Do you mean this as a single piece of data, which is what the singular form means?'' {\\em ``statistic'' is used here as: ``a function of a sample where the function itself is independent of the sample's distribution.'' \\url{http://en.wikipedia.org/wiki/Statistic}}} used for finding matched circles exists for a circle radius $\\sim 11\\ddeg$ and a $-36\\ddeg$ twist, but it is not statistically significant. \\nocite{Aurich2005circ}{Aurich} {et~al.} (2006) also made a circles analysis of the WMAP first-year data, using their own estimator and weight function, and, in contrast with \\nocite{RLCMB04}{Roukema} {et~al.} (2004), \\nocite{KeyCSS06}{Key} {et~al.} (2007), and \\nocite{LewRouk2008}{Lew} \\& {Roukema} (2008), did not find any signal at $11\\ddeg$. On the other hand, they did find a tentative PDS signal at $\\Omtot \\approx 1.015$, or equivalently, a circle radius $\\alpha \\approx 40\\ddeg$. The signal is weaker than they expected, but the authors note that uncertainties due to foregrounds and noise structures in the data make it premature to draw firm conclusions. A disadvantage of the identified circles approach is that it is based on the information in a relatively small number of points on the sky map of temperature fluctuations, making it sensitive to small errors in the data or analysis and requiring prohibitively long computations. Is it \\langed{No, the word ``even'' would be wrong here.} possible to generalise from the identified circles principle? Moreover, leaving aside the debate about matched circles statistics, observably multiply connected models could reasonably be said to have satisfied only one prediction so far, that of a cutoff in the density fluctuation spectrum on a large scale. Can predictions of the PDS model itself be tested, for example, using the identified circles principle or an extension of it? In searches for matched circles, the statistic used is usually some variation on what can be considered to be the value of the two-point cross-correlation function of observed temperature fluctuations at pairs of points in the covering space, $\\ximc(r)$, where the two points lie on different copies of the SLS, and the separation of the two points in the pair (on the different copies of the SLS) is zero in the covering space. \\postrefereechanges{ We can write this function as if it were possible to sample temperature fluctuations as point objects at arbitrary spatial points throughout the covering space, i.e. including arbitrarily low redshifts, as well as points beyond the SLS: \\begin{eqnarray} \\ximc(r) &\\equiv& \\left< \\delta T (x_{i_1}) \\delta T \\left[g_j({x_{i_2}})\\right] \\right>_{i_1,i_2,j} % \\label{e-ximc-defn} \\end{eqnarray} averaging over triples $(i_1,i_2,j)$ satisfying $d\\left( x_{i_1}, \\left[g_j({x_{i_2}})\\right] \\right) = r$ and $g_j\\not= I$ (the identity $I$ is removed because we only want the cross-correlation). Here, $x_{i_k}$ for $k=1,2$ are arbitrary points on the SLS considered to be located at their comoving spatial positions, $g_j$ for $j=1, \\ldots, 12$ are the 12 holonomy transformations in the binary icosahedral group $\\Gamma = I^*$ operating on the covering space $S^3$ that match opposite faces of the fundamental polyhedron of the PDS, $d(x,y)$ is the comoving spatial geodesic distance between two points $x,y$ in the comoving covering space (i.e. an arc-length on the covering space $S^3$ embedded in $\\mathbb{R}^4$, e.g. \\nocite{Rouk01-4D}{Roukema} 2001). The temperature fluctuations on one copy of the SLS, $\\delta T(x)$, are extended to the whole covering space via the holonomy transformations, i.e., \\begin{equation} \\forall g_j \\in \\Gamma, \\;\\; \\delta T(x) = \\delta T \\left[g_j(x)\\right] . \\label{e-temp-covspace} \\end{equation} The latter assumption, i.e. Eq.~(\\ref{e-temp-covspace}), enables rewriting Eq.~(\\ref{e-ximc-defn}) in a way that becomes observationally realistic, \\begin{eqnarray} \\ximc(r) &\\equiv& \\left< \\delta T (x_{i_1}) \\delta T ({x_{i_2}}) \\right>_{i_1,i_2,j}, % \\label{e-ximc-two} \\end{eqnarray} again subject to the conditions that the average is taken over pairs satisfying $d\\left( x_{i_1}, \\left[g_j({x_{i_2}})\\right] \\right) = r$ and $g_j\\not= I$. The auto-correlation function can then be written \\begin{eqnarray} \\xisc(r) \\equiv \\left< \\delta T (x_{i_1}) \\delta T ({x_{i_2}}) \\right>_{i_1,i_2} , \\label{e-xisc-defn} \\end{eqnarray} averaging over pairs satisfying $ d\\left( x_{i_1}, x_{i_2} \\right) = r$. In words, $\\xisc$ is the 3-spatial auto-correlation function for pairs of points on a single copy of the SLS, while $\\ximc$ is the 3-spatial cross-correlation function for pairs of points where the two members lie on different copies of the SLS in the covering space. Now rewrite \\langed{$S$ was ``written'' in Roukema et al. 2004, so now it is {\\bf re}written.} the statistic $S$ used in matched circles searches as follows [see e.g. Eq.~(9) of \\nocite{RLCMB04}{Roukema} {et~al.} (2004), with the normalisations (denominator and monopole $T$) ignored for simplicity], \\begin{eqnarray} S &\\equiv& { \\left< {\\delta T }_i \\; {\\delta T }_j \\right>_{(i,j)} } \\label{e-s-rlcmb04} \\nonumber \\\\ & = & \\ximc (0) \\label{e-ximc-zero} \\end{eqnarray} for the case where pairs of points $(i,j)$ are (hypothetically) multiply-imaged locations located on a pair of matching circles in the covering space. We use $r=0$ here because that is the defining characteristic of matched circles --- a pair of matching points is a match because the two points are the same space-time points, i.e. they are separated by $r=0$ when their topological images located on adjacent copies of the SLS in the covering space are considered. } \\newcommand\\rmax{2} Here, generalise from the zero separation cross-correlation $\\ximc(0)$ to cross-correlations at larger separations $r>0$. We can expect that \\begin{equation} \\ximc(r > 0) < \\ximc(0), \\end{equation} i.e. that correlations weaken with separation and that, in general, $\\ximc(r)$ should be approximately equal to the auto-correlation function $\\xisc(r)$ on scales much smaller than \\langed{$\\ll$ and $<$ differ.} the ``size'' of the fundamental domain, e.g. the in-diameter $2r_{-}$ \\nocite{LR99}(e.g. fig~10, {Luminet} \\& {Roukema} 1999) \\begin{eqnarray} \\ximc(r > 0) \\sim \\xisc(r > 0) & &, \\mbox{ if } r \\ll 2r_{-}, \\end{eqnarray} since there is no statistical, physical distinction between a pair of points on different copies of the SLS and a pair of points on a single copy of the SLS, apart from effects that are not locally isotropic, such as the Doppler effect. Since $\\xisc$ is generally large at small $r$ and small at large $r$, obtaining a large \\langed{Using ``strong'' would be confusing, since we use ``large'' a few words earlier in the same sentence. Moreover, ``strong'' has connotations either of physical force or human force, which is not a useful connotation here. ``Close'' would be utterly confusing to the reader.} correlation requires using a range of length scales $r$ that are relatively small, e.g., \\begin{eqnarray} \\ximc(r \\ltapprox \\rmax \\hGpc) &\\sim& \\xisc(r \\ltapprox \\rmax \\hGpc) \\label{e-ximc-pdsgood} \\end{eqnarray} should be a relatively high positive value if the multiply connected model being studied is the correct model. Another way of saying this is that this test compares the spatial two-point cross-correlation function of mapped (in the sense of the holonomy transformation $g$, which maps one copy of a point in the covering space to one of its images) and unmapped temperature fluctuations in the covering space, where the two points in any pair are on different copies of the SLS, against the auto-correlation function on a single copy of the SLS. To see yet another way of thinking about this, suppose that the PDS model at a given orientation and implied circle size is correct and that temperature fluctuations occur as point objects, each emitting isotropically. The covering space {\\em cross}-correlation function, based on pairs of points that are {\\em close to one another due to the application of one or more holonomy transformations $g$, but are not close to one another on a single copy of the SLS}, should then be statistically equivalent (apart from the Doppler effect and foreground/projection effects) to a sampling from the spatial {\\rm auto}-correlation function $\\xisc$, which can be estimated using points that lie on a single copy of the SLS. {\\em This gives us our first prediction to test the PDS hypothesis: Eq.~(\\ref{e-ximc-pdsgood}) should only be expected to hold if the physically correct PDS (or other) 3-manifold is assumed and is modelled at its correct astronomical orientation.} If the PDS model is correct but a wrong orientation is chosen, then the cross-correlation $\\ximc$ inferred from it on small length scales --- in the covering space since it is the cross-correlation --- should represent correlations of pairs of points that in reality are {\\em not} close together, due to the erroneous orientation that causes mappings from one copy of the SLS to another to be incorrect. Since the correlation of distant points is, in general, small, we have for this incorrect PDS model: \\begin{eqnarray} \\ximc(r \\ltapprox \\rmax \\hGpc) &\\sim& \\xisc(r \\gg \\rmax \\hGpc) \\nonumber \\\\ & \\ll & \\xisc(r \\ltapprox \\rmax\\hGpc) \\label{e-ximc-pdsbad-first} \\end{eqnarray} i.e. \\begin{eqnarray} \\ximc(r \\ltapprox \\rmax \\hGpc) & \\ll & \\xisc(r \\ltapprox \\rmax\\hGpc), \\label{e-ximc-pdsbad} \\end{eqnarray} in contrast to Eq.~(\\ref{e-ximc-pdsgood}). Similarly, if the PDS model is incorrect, then arbitrary orientations should also yield small cross-correlations on small length scales, as in Eq.~(\\ref{e-ximc-pdsbad}). In this paper, the initial aim is to maximise $\\ximc(r \\ltapprox \\rmax \\hGpc)$ relative to $\\xisc(r \\ltapprox \\rmax \\hGpc)$, varying PDS models over the parameter space of different orientations and circle sizes. (In the actual calculations, we used a range of scales below and a little above $\\sim \\rmax \\hGpc$: see \\SSS\\ref{s-scales} and Eq.~(\\ref{e-range-Gpc}) for details.) This should lead to an estimate of the best orientation and circle size for the PDS model, given the observational data. However, to further test the PDS model, consider a ``generalisation'' from the mathematically correct PDS model to a class of pseudo-models for which the cross-correlation function can be calculated, but for which most\\footnote{The mathematical term ``almost all'' could be used here, but when taking observational and numerical calculation uncertainties into account, the word ``most'' is more realistic.} of the members of the class are physically invalid. Maximising $\\ximc(r \\ltapprox \\rmax \\hGpc)$ (relative to $\\xisc(r \\ltapprox \\rmax \\hGpc)$) over the extended parameter space defined by this class should then have little chance of yielding an optimal model that is valid, unless the PDS model is astronomically correct. The ``generalisation'' that we define here is to allow the ``twist'' angle $\\phi$ to be arbitrary in $\\left[0,2\\pi\\right]$. The twist angle can be described as follows. In a single-action, spherical 3-manifold \\nocite{GausSph01}(e.g., Sect. 4.1, {Gausmann} {et~al.} 2001) thought of as embedded in 4-dimensional Euclidean space, $\\mathbb{R}^4$, any holonomy transformation is a Clifford translation that rotates in $\\mathbb{R}^4$ about the centre of the hypersphere in one 2-plane by an angle of $\\pi/5$, and also by $\\pi/5$ in an orthogonal 2-plane. From a close-up perspective of the SLS rather than looking at the whole hypersphere $S^3$, or in other words, from a projection into $\\mathbb{R}^3$, one 4-rotation can be thought of as a {\\em translation} in $\\mathbb{R}^3$ from one circle in a matched circle pair to the other member of the pair on the opposite side of the SLS, and the second 4-rotation is then thought of as a {\\em twist} around a vector in the translation direction joining the centres of the two identified circles. This full motion is frequently termed a ``screw motion''. The ``twist'' is the second of these two rotation angles. Physically, the only two possible twist angles are $\\pm \\pi/5 .$ If the cross-correlation were to be calculated by filling one copy of the fundamental domain with a uniform distribution of points and then mapping this set of points to copies of the fundamental domain in the covering space, then for an invalid twist angle, some regions of space near the first copy would have either zero, two, three, or more times the density of points in the first copy of the fundamental domain. In other words, sharp discontinuities would occur in calculating the cross-correlation. However, a practical way of estimating the cross-correlation is the method represented in Eq.~(\\ref{e-s-rlcmb04}), i.e. multiplying temperature fluctuations and then averaging them. Using this method rather than multiplying densities of points, the effect mentioned in the previous paragraph should only lead to slightly more or less frequent sampling in some regions of the cross product of the covering space with itself, not to a modification of the correlations themselves. The continuous nature of this method, as opposed to the discrete nature of starting with a uniform distribution of points in the fundamental domain, can be seen as follows. For a given pair of observed locations $(x_i, x_j)$ on one copy of the SLS, their comoving spatial separation $d[x_i,g(x_j,\\phi)]$ for a given ``generalised'' \\postrefereechanges{ isometry } $g$, which depends on the generalised twist $\\phi$, changes continuously through the values $\\phi = \\pm \\pi/5,$ not discretely. \\langed{Either ``and'' or ``but'' would be wrong. ``i.e.'' would be correct but would sound pedantic.} Moreover, the product of the observed temperature fluctuations remains constant. Intuitively, we could say that as $\\phi$ varies, the geodesics formed by pairs $[x_i,g_1(x_j,\\phi)]$ do not ``know'' when they meet other pairs $[x_i,g_2(x_j,\\phi)]$, where $g_1$ and $g_2$ are two holonomy transformations mapping the fundamental domain to two copies of the fundamental domain adjacent to one another in the covering space. \\langed{This should be clearer.} If the PDS model is correct, then calculating $\\ximc(r)$ using temperature fluctuations and applying isometries using an arbitrary twist angle should give $\\ximc (r \\ltapprox \\rmax \\hGpc)\\ll \\xisc (r \\ltapprox \\rmax \\hGpc)$ as in Eq.~(\\ref{e-ximc-pdsbad}), but give $\\ximc (r \\ltapprox \\rmax \\hGpc)\\sim \\xisc (r \\ltapprox \\rmax \\hGpc)$ as in Eq.~(\\ref{e-ximc-pdsgood}) when the twist angle is correct. {\\em This gives the second prediction of the PDS model. The maximal cross-correlation for this generalised PDS model estimated using correlations of temperature fluctuations should not only exist as a robust maximum, but it should also give a twist angle of either $\\pm \\pi/5$.} If the simply connected, perfectly flat model is correct, then the PDS model is incorrect, and it should either be difficult to find a robust maximal correlation $\\ximc (r \\ltapprox \\rmax \\hGpc)$ in this extended 5-parameter space, or else an arbitrary twist angle $\\phi$ should result, with only a small chance of $\\phi$ being close to either of the two values expected for the PDS. An estimate of how close an ``arbitrary'' twist angle should lie to one of the two PDS values can be made as follows. Given the assumption that \\begin{eqnarray} \\xisc(r \\gg \\rmax \\hGpc) & \\ll & \\xisc(r \\ltapprox \\rmax\\hGpc), \\label{e-xisc-zeroatbigr} \\end{eqnarray} we can expect that this arbitrary angle should be selected from a uniform probability density distribution on $\\left[0,2\\pi\\right]$. In principle, depending on the assumptions made about the complex statistical properties of the WMAP temperature fluctuation maps, it could be possible for this distribution to be non-uniform, even for a non-PDS space model. However, the estimated spatial auto-correlation function $\\xisc$ for $r \\gtapprox 10${\\hGpc} is close to zero, as shown in fig~16 of \\nocite{WMAPSpergel}{Spergel} {et~al.} (2003) as a function of angular separation on the SLS ($S^2$), and explicitly as a function of spatial separation in Fig.~\\ref{f-auto} here, so the assumption does appear to be supported by the \\langed{The only serious direct empirical evidence is that of WMAP, hence ``the'' rather than implicit ``some''.} empirical evidence. For the uniform distribution assumption, let us define \\begin{eqnarray} \\Delta\\phi &\\equiv & \\min \\left( \\left|\\phi - \\frac{\\pi}{5}\\right|, \\left|\\phi - \\frac{9\\pi}{5}\\right| \\right) \\label{e-defnDphi} \\end{eqnarray} for $\\phi \\in \\left[0,2\\pi\\right]$. The chance of the observational optimal phase $\\phiWMAP$ being close to $\\pi/5$ or $9\\pi/5$ is then \\begin{eqnarray} P\\left( \\Delta\\phi < \\Delta\\phiWMAP \\right) &=& \\left\\{ \\begin{array}{l l} 2 \\frac{\\Delta\\phi}{\\pi} , & \\mbox{ if } \\Delta\\phi \\le \\frac{\\pi}{5} \\\\ \\frac{1}{5} + \\frac{\\Delta\\phi}{\\pi} , & \\mbox{ if } \\frac{\\pi}{5} \\le \\Delta\\phi \\le \\frac{4\\pi}{5} . \\\\ \\end{array} \\right. \\label{e-probphi} \\end{eqnarray} The piecewise nature of this function is because; e.g. there are four ways in which $\\phi$ can differ from $\\pm \\pi/5$ by a small angle such as $10\\ddeg$, but only two ways in which it can differ from $\\pm \\pi/5$ by a large angle such as $100\\ddeg$. To search for an optimal solution in the parameter space, a Metropolis-Hastings version of a Markov-chain Monte Carlo (MCMC) method is used \\nocite{Neal93}(e.g., Sect. 4.2, {Neal} 1993, and references therein) over the five-dimensional parameter space \\begin{equation} \\{ (l,b,\\theta,\\alpha,\\phi) \\}, \\label{e-parameterspace} \\end{equation} where the parameters represent dodecahedron orientation ($l$, $b$, $\\theta$), circle size $\\alpha$ and twist phase $\\phi$. All parameters are initialised as arbitrary angles, except that the circle size is constrained to $5\\ddeg \\le \\alpha \\le 60\\ddeg$ both initially and during the whole chain. See \\SSS\\ref{s-mcmc-method} for more details. We also ran some MCMC chains starting with the parameters of the \\nocite{RLCMB04}{Roukema} {et~al.} (2004) ``hint'' of a PDS solution with circles of size $\\sim 11 \\ddeg$. If this solution is correct, then we would expect the chains to remain localised around that same solution. If the solution is wrong and if a correct solution exists, then the chains should move towards that correct solution, even though starting at a wrong point. If the solution is wrong and if no correct solution exists, then we would expect the chains to move randomly and fail to find a strong maximum. The correlation function definitions are described in \\SSS\\ref{s-corr-method}, the probability estimator used to compare multiple-SLS cross-correlation and single-SLS auto-correlation functions is presented in \\SSS\\ref{s-prob-method}, and the MCMC method is described in \\SSS\\ref{s-mcmc-method}. Results are presented in \\SSS\\ref{s-results}. Discussion follows in \\SSS\\ref{s-disc} and conclusions are made in \\SSS\\ref{s-conclu}. For references on cosmic topology in general, please see the first known article on the subject \\nocite{Schw00,Schw98}({Schwarzschild} 1900, 1998), a short beginner's review of cosmic topology \\nocite{Rouk00BASI}({Roukema} 2000), more in-depth reviews \\nocite{LaLu95,Lum98,Stark98,LR99}({Lachi\\`eze-Rey} \\& {Luminet} 1995; {Luminet} 1998; {Starkman} 1998; {Luminet} \\& {Roukema} 1999), workshop proceedings \\nocite{Stark98,BR99}({Starkman} 1998; {Blanl{\\oe}il} \\& {Roukema} 2000), and lists of two-dimensional and three-dimensional methods (Table~2, \\nocite{LR99}{Luminet} \\& {Roukema} 1999; \\nocite{ULL99b,Rouk02topclass,RG04}{Uzan} {et~al.} 1999; {Roukema} 2002; {Rebou\\c{c}as} \\& {Gomero} 2004). For background on spherical, multiply connected spaces, see \\nocite{Weeks2001}{Weeks} (2001), \\nocite{GausSph01}{Gausmann} {et~al.} (2001), \\nocite{LehSph02}{Lehoucq} {et~al.} (2002), \\nocite{RiazSph03}{Riazuelo} {et~al.} (2004), and \\nocite{LumNat03}{Luminet} {et~al.} (2003). For the identified circles principle, of which the present method can be thought of as an extension, see \\nocite{Corn96,Corn98b}{Cornish} {et~al.} (1996, 1998). One correction to a point made in much of the above literature concerns the independence of the metric and global topology. For example, \\nocite{LR99}{Luminet} \\& {Roukema} (1999) wrote, ``However, general relativity deals only with local geometrical properties of the universe, such as its curvature, not with its global characteristics, namely its topology''. The error here is that global topology {\\em can}, at least in certain cases, generate a local effect, and thus affect local geometrical properties. See the $T^3$, Newtonian weak field limit, heuristic calculation in \\nocite{RBBSJ06}{Roukema} {et~al.} (2007): if the Universe contains inhomogeneities and is not expanding perfectly isotropically, then at least in the $T^3$ case, \\postrefereechanges{ additional accelerations and decelerations exist in the different fundamental directions, in such a way that they tend to equalise the slightly different expansion rates.} Comoving coordinates are used when discussing distances (i.e. ``proper distances'', \\nocite{Wein72}{Weinberg} 1972, equivalent to ``conformal time'' if $c=1$). We write the Hubble constant as $H_0 \\equiv 100 h$\\kms/Mpc. ", "conclusions": "\\label{s-conclu} It seems hard to avoid the conclusion that cross-correlations of temperature fluctuations on would-be adjacent copies of the SLS, which are distant from one another and are (on average) very weakly correlated according to the WMAP 3-year observations, (i) imply a highly cross-correlated ``generalised'' Poincar\\'e dodecahedral space symmetry (Figs~\\ref{f-ilc_lbth_N}--\\ref{f-toh_lbth_N}, Table~\\ref{t-dodec}) at which these on-average uncorrelated fluctuations happen to be well-correlated with one another (Fig.~\\ref{f-cross}), and, moreover, (ii) this favoured solution is dominated by a signal whose twist phase $\\phi$ lies within a few degrees of one of the two twist phases necessary for a valid PDS model (Table~\\ref{t-alpha-phi}). These two successful predictions of the PDS model follow the WMAP confirmation of the generic prediction of small universe models, a power cutoff in structure statistics on large scales, and the solution appears to be consistent with quite different PDS analyses of the WMAP data. Do we really live in a Poincar\\'e dodecahedral space? Further constraints either for or against the model are certainly still needed, but the evidence in favour of a PDS-like signal in the WMAP data does seem to be accumulating." }, "0801/0801.2003_arXiv.txt": { "abstract": "This is the first paper of a series focused on investigating the star formation and evolutionary history of the two early-type galaxies NGC~1407 and NGC~1400. They are the two brightest galaxies of the NGC~1407 (or Eridanus-A) group, one of the 60 groups studied as part of the Group Evolution Multi-wavelength Study (GEMS). Here we present new high signal-to-noise long-slit spectroscopic data obtained at the ESO 3.6m telescope and high-resolution multi-band imaging data from the HST/ACS and wide-field imaging from Subaru Suprime-Cam. We spatially resolved integrated spectra out to $\\sim$~0.6 (NGC~1407) and $\\sim$~1.3 (NGC~1400) effective radii. The radial profiles of the kinematic parameters $v_{rot}$, $\\sigma$, $h_{3}$ and $h_{4}$ are measured. The surface brightness profiles are fitted to different galaxy light models and the colour distributions analysed. The multi-band images are modelled to derive isophotal shape parameters and residual galaxy images. The parameters from the surface brightness profile fitting are used to estimate the mass of the possible central supermassive black hole in NGC~1407. The galaxies are found to be rotationally supported and to have a flat core in the surface brightness profiles. Elliptical isophotes are observed at all radii and no fine structures are detected in the residual galaxy images. From our results we can also discard a possible interaction between NGC~1400, NGC~1407 and the group intergalactic medium. We estimate a mass of $\\sim$1.03~$\\times$~10$^{9}$~$M_{\\sun}$ for the supermassive black hole in NGC~1407 galaxy. ", "introduction": "Early-type galaxies are estimated to contribute to at least half, and perhaps as much as three quarters, of the stellar mass in the local Universe \\citep{bell03}. Understanding their formation and evolution is therefore a fundamental objective of astrophysical research. \\begin{table*} \\begin{center} \\begin{tabular}{ccccccccccc} \\hline \\hline Name & R.A. & DEC. & Type &PA & v & $\\sigma_{0}$ &$r_{e}$ & M$_{B}$ & L$_{X}$\\\\ &(J2000) & (J2000) & &(degrees) &(km s$^{-1}$)& (km s$^{-1}$)& (arcsec) &(mag) & (erg~s$^{-1}$) \\\\ \\hline NGC~1407 & 03:40:11.9 & $-$18:34:49 & E0 &35 &1779 $\\pm$ 9 & 305.0 $\\pm$ 10.0 &72$\\arcsec$ & $-$21.22 & 1.00~$\\times$~10$^{41}$ \\\\ NGC~1400 & 03:39:30.8 & $-$18:41:17 & E$-$S0 & 40 & 558 $\\pm$ 14 & 279.0 $\\pm$ 2.0 &27$\\arcsec$ & $-$19.95 & 1.32~$\\times$~10$^{40}$ \\\\ \\hline \\hline \\end{tabular} \\end{center} \\caption{Galaxies properties. Type, P.A.: morphological type and position angle of the major axis from HyperLEDA; v, $\\sigma_{0}$, $r_{e}$: velocity, central velocity dispersion and effective radius from HyperLEDA; M$_{B}$: absolute B-band magnitude obtained from HyperLEDA; L$_{X}$: X-ray luminosity from \\citep{osullivan01}.} \\label{ga_prop} \\end{table*} The debate on whether early-type galaxies are formed by monolithic collapse at high redshifts or via hierarchical merging is still not settled. Photometrically, the tightness of the colour-magnitude diagram suggests that the bulk of the stars were formed at high redshifts, z~$>$~1 (e.g \\citealt{bower92}; \\citealt{kodama99}). However, signs of recent mergers and interactions are often present when images of early-type galaxies are examined (e.g. \\citealt{ft92}). Shells and ripples are believed to be formed by mergers (e.g. \\citealt{quinn84}) or by interactions (e.g. \\citealt{thomson90}; \\citealt{thomson91}). Deviations from perfect elliptical isophotes are correlated with the galaxy shape and amount of rotation support (e.g. \\citealt{khoc05}). Furthermore, boxyness is often associated with X-ray and radio activities which may indicate an evolutionary history more affected by mergers (e.g. \\citealt{bender89}). Hence the detailed photometric study of early-type galaxies may yield important clues on their past merger activities. Clues to the formation of early-type galaxies may also come from their kinematics. Studies on the global kinematics have established that elliptical galaxies as a class are supported by velocity anisotropy (e.g. \\citealt{binney76}; \\citealt{binney78}). Detailed kinematic studies often reveal kinematic distinct cores (e.g. \\citealt{emsellem04}) and non-relaxed structure (e.g. \\citealt{balce99}) which may be related to the way the galaxy is formed and to the merger history. It is possible to measure the shape of absorption lines, hence the Line-Of-Sight Velocity Distribution (LOSVD), which will tell us about the velocity anisotropy and hence constrain the orbital families (\\citealt{bn90}; \\citealt{rix92}; \\citealt{van93}). The LOSVD is often parametrised by the mean velocity $v_{rot}$ and velocity dispersion $\\sigma$, plus higher order moments ($h_{3}$, $h_{4}$ ....) of a Gauss-Hermite series. The $h_{3}$ and $h_{4}$ offer extra information on the asymmetric and symmetric deviation, respectively, away from a perfect Gaussian. Observationally, we now have the ability to obtain spatially resolved high signal-to-noise integrated spectra out to large galactic radii and hence larger mass fractions. The high-resolution multi-band imaging data provide further insights on the radial distribution of the galaxy light and allow meaningful comparisons with spectroscopically derived results. Therefore, we are able to consider physical mechanisms acting locally and to investigate how their properties vary with the galactocentric radius. In this series of papers we focus on the internal properties of the two early-type galaxies NGC~1407 and NGC~1400 in an effort to understand their star formation and evolutionary history. In this first paper (hereafter Paper~I), we present spatially resolved radial profiles of the kinematic parameters $v_{rot}$, $\\sigma$, $h_{3}$ and $h_{4}$ out to $\\sim$~0.6 (NGC~1407) and $\\sim$~1.3 times (NGC~1400) the galaxies' effective radii ($r_{e}$; the radius within which half the galaxy light is contained). We also present spatially resolved multi-band high-resolution wide-field imaging data out to $\\sim$~1.4$r_{e}$ for both galaxies. The spatial distributions of the galaxy light profiles are analysed and the isophotal ellipticity radial profiles and deviation parameters recovered. The residual images of the two galaxies are inspected for fine structure (e.g. dust, ripples, tidal tails, boxy or disky structure, etc.). In the second paper (\\citealt{spola07b}, hereafter Paper~II) we will present spatially resolved stellar population parameters from the same high-S/N long-slit spectroscopic data used here. The aim of Paper~II will be to interpret and combine the results from the stellar populations analysis with the kinematic and photometry results. This paper is organised as follows. In Section~2 we describe the two sample galaxies. In Sections~3 we describe the spectroscopic and photometric observations, together with an explanation of the relevant data reduction procedures. Section~4 presents the spatially resolved radial kinematic profiles. Section~5 presents the spatially resolved surface photometry results. In Section~6 we discuss a possible interaction between NGC~1400 and NGC~1407 and the group intergalactic medium (IGM). In Section~7 we summarise our results. ", "conclusions": "This is the first paper of a series with the aim of studying the star formation and evolutionary history of the two early-type galaxies NGC~1407 and NGC~1400. Here, we have presented spatially resolved radial kinematics and surface photometry. The spectroscopic analysis is performed using high signal-to-noise long-slit data obtained at the ESO 3.6m telescope. The imaging study is based on high-resolution multi-band data from the HST/ACS and wide-field imaging from Subaru Suprime-Cam. The high-quality of the data allowed us to focus our study on the properties of spatially resolved galactic regions and to trace the overall radial trends. The radial profiles of the kinematic parameters $v$, $\\sigma$, $h_{3}$ and $h_{4}$ have been recovered. From the imaging analysis we obtained surface brightness and colour index profiles. We also measured the radial profiles of the isophotal shape parameters and created residual galaxy images to search for fine structure. \\begin{list}{}{} \\item \\textbf{NGC~1407.} The kinematic study suggests a rotationally supported galaxy (or marginally anisotropic) with the presence of a possible kinematically decoupled core; the detection is uncertain and potentially may be caused by a misalignment of the slit with the nucleus. The surface brightness profiles reveal a flat core probably caused by the presence of a central supermassive black hole, with an estimated mass of $\\sim$1.03~$\\times$~10$^{9}$~$M_{\\sun}$. We find NGC~1407 to be an elliptical galaxy (A$_{4}$=0) with a mean position angle of 59$^{\\circ}$ and a small ellipticity value of 0.05. No fine structure is detected in the residual images. \\item \\textbf{NGC~1400.} The anisotropy parameter and the kinematic profiles indicate NGC~1400 to be rotationally supported with evident minor axis rotation and flattening due to fast rotation. The galaxy was already found to have a flat core in the surface brightness profile (\\citealt{lauer95}). The galaxy isophote modelling shows elliptical isophotes at all radii (A$_{4}$=0) and no fine structure is detected in the residual galaxy images. The ellipticity is more pronounced, 0.11, and the position angle is 36$^{\\circ}$. From our results we would tend to classify NGC~1400 as an E1 galaxy, in contrast with the E$-$S0 morphological classification proposed in literature. \\end{list} We found no evidence to support an interaction between NGC~1400 and NGC~1407 in our data. We speculate that what we are now witnessing might be the first infall of a subgroup, dominated by the NGC~1400 galaxy, into the NGC~1407 group." }, "0801/0801.2529_arXiv.txt": { "abstract": "{ In this paper we present a study of chemical abundances in six star-forming regions. Stellar parameters and metallicities are derived using high-resolution, high S/N spectra of weak-line T-Tauri stars in each region. The results show that nearby star-forming regions have a very small abundance dispersion (only 0.033\\,dex in [Fe/H]). The average metallicity found is slightly below that of the Sun, although compatible with solar once the errors are taken into account. The derived abundances for Si and Ni show that the observed stars have the abundances typical of Galactic thin disk stars of the same metallicity. The impact of these observations is briefly discussed in the context of the Galactic chemical evolution, local inter-stellar medium abundances, and in the origin of metal-rich stars in the solar neighbourhood (namely, stars more likely to harbour planets). The implication for future planet-search programmes around very young, nearby stars is also discussed. ", "introduction": "As of November 2007, about 250 extra-solar planets have been discovered orbiting solar-type stars\\footnote{For a continuously updated list see tables at http://www.exoplanets.eu or http://www.exoplanet.eu}, most of which were detected thanks to the development of high precision radial-velocity instruments \\citep[for a review see][]{Udry-2007}. Complementary high accuracy spectroscopic studies have shown that stars hosting giant planets are particularly metal-rich when compared with single field stars \\citep[e.g.][]{Gonzalez-1997, Gonzalez-2001,Santos-2001,Santos-2004b,Santos-2005a, Fischer-2005}. This observation is helping to better understand the processes of planet formation and evolution. The high metal content of some of the planet-hosts has raised a number of interesting questions regarding their origin. It has been proposed that the high metallicities could be the result of the in-fall of planetary (metal-rich) material into the stellar outer convective envelope \\citep[e.g.][]{Gonzalez-1998,Murray-1998}, although current results do not support this hypothesis \\citep[][]{Santos-2004b,Santos-2005a,Fischer-2005}. The excess metallicity observed likely reflects the higher (average) metal abundances of the clouds of gas and dust that originated the star and planet systems. Such conclusions are also supported by the most recent planet formation models \\citep[][]{Ida-2004b,Benz-2006}. Because the large majority of planet hosting stars are solar-neighbourhood thin-disk objects \\citep[e.g.][]{Ecuvillon-2007}, one would expect nearby T-Tauri stars and star-forming regions [{\\sc sfr}s] that are metal-rich to exist. Their existence is expected if a well-defined age-metallicity relation exists in the solar-neighbourhood \\citep[][]{Pont-2004b,Haywood-2006}. Previous studies \\citep[][]{Padgett-1996,James-2006}, however, have not succeeded in finding evidence of high metallicity, T-Tauri stars in nearby {\\sc sfr}s. Whether this is due to the absence of such regions, or to the relatively small number of stars and {\\sc sfr}s observed was not clear. Considering the strong metallicity-giant planet connection, the detection of nearby, metal-rich {\\sc sfr}s would provide preferential targets for future planet searches around young T-Tauri stars. The detection of planets orbiting T-Tauri stars would provide observational constraints of paramount importance for the study of planet formation scenarios, since it gives us some hint about the timescales for planetary formation. In this paper, we continue the metallicity programme started in \\citet[][]{James-2006}, by increasing the number of observed young T-Tauri stars in each {\\sc sfr}, and by adding abundance determinations for objects that have not yet been studied. In all, we present accurate stellar parameters and chemical abundances for 19 weak line T-Tauri stars [{\\sc wtts}s] in six nearby {\\sc sfr}s, namely Chamaeleon, Corona Australis, Lupus, Rho-Ophiuchus, Taurus and the Orion Nebula Cloud. The results of this study are presented and their implications discussed. ", "conclusions": "In this paper, we present stellar metallicities for six nearby {\\sc sfr}s, derived from high resolution spectroscopy of young {\\sc wtts}s. The results show that all the studied {\\sc sfr}s have metallicities between $-$0.13 and $-$0.05\\,dex, with a very small scatter of 0.033\\,dex around the mean value of $<$[Fe/H]$>$=$-$0.08\\,dex (compatible to solar within the uncertainties). The analysis of the abundances of silicon (alpha-element) and nickel (iron-peak element) also suggests that the abundance ratios in the studied {\\sc sfr}s are typical of those found in solar neighbourhood thin disk stars of similar metallicity. Together with other results from the literature, these observations suggest that the chemical abundances in the nearby {\\sc ism} are extremely uniform and not strongly above solar. Such a conclusion may have important implications for the models of Galactic evolution. The results presented here have also important implications on planet searches around young (active) stars. Although these searches are severely limited in current radial-velocity surveys \\citep[e.g.][]{Saar-1997,Santos-2000a,Paulson-2002}\\footnote{See however recent detection by \\citet[][]{Setiawan-2008}}, several instruments are now being built that will change this situation. In particular, new generation adaptive optics systems like the {\\sc eso sphere} planet finder or the Gemini {\\sc gpi} project, as well as a new generation of interferometric instruments like {\\sc prima} ({\\sc eso}), will allow us to study, in unprecedented detail, the existence of long period planets around young solar-type stars. The rarity of metal-rich nearby star-forming regions may limit the goals of these projects if the metallicity-giant planet connection is still present for systems with long orbital periods. In this context it will also be very interesting to study in detail the abundances of nearby young, post-T Tauri stars \\citep[e.g.][]{Zuckermann-2004,Torres-2006}. Such studies would also help to settle the question about the metallicity of the local {\\sc ism} \\citep[e.g.][]{Sofia-2001}. If stars with planets have originated from inner galactic regions, it may be interesting to understand how planetary evolution (and survival) may suffer with stellar orbital diffusion. This conclusion may further have impact on the understanding of the frequency of planets in the Galaxy. As a very important by-product, the observations presented in this paper can further be used to investigate the effects of the metallicity on the properties of stellar populations such as rotation, multiplicity degree, or magnetic activity. These studies will be presented in a separate paper, James et al. (in prep.)." }, "0801/0801.4413_arXiv.txt": { "abstract": "We investigate the tidal interaction between a low-mass planet and a self-gravitating protoplanetary disk, by means of two-dimensional hydrodynamic simulations. We first show that considering a planet freely migrating in a disk without self-gravity leads to a significant overestimate of the migration rate. The overestimate can reach a factor of two for a disk having three times the surface density of the minimum mass solar nebula. Unbiased drift rates may be obtained only by considering a planet and a disk orbiting within the same gravitational potential. In a second part, the disk self-gravity is taken into account. We confirm that the disk gravity enhances the differential Lindblad torque with respect to the situation where neither the planet nor the disk feels the disk gravity. This enhancement only depends on the Toomre parameter at the planet location. It is typically one order of magnitude smaller than the spurious one induced by assuming a planet migrating in a disk without self-gravity. We confirm that the torque enhancement due to the disk gravity can be entirely accounted for by a shift of Lindblad resonances, and can be reproduced by the use of an anisotropic pressure tensor. We do not find any significant impact of the disk gravity on the corotation torque. ", "introduction": "$} \\label{sec:appA} In this section, we give the expressions of the radial and azimuthal self-gravitating accelerations $g_r$ and $g_{\\varphi}$, smoothed over the softening length $\\varepsilon_{\\rm sg}$. We use the variables ($u=\\log(r/r_{\\rm min})$, $\\varphi$), where $r_{\\rm min}$ denotes the inner edge radius of the grid. With this set of coordinates, $g_r (u,\\varphi)$ reads \\begin{eqnarray} \\left. g_r (u, \\varphi) \\right. = &-&Ge^{-u/2}\\,\\int_{0}^{u_{\\rm max}} \\int_0^{2\\pi} S_r(u^{'}, \\varphi^{'})\\,K_r(u-u^{'},\\varphi-\\varphi^{'})\\,du^{'}d\\varphi^{'}\\nonumber\\\\ &+&G\\Sigma(u,\\varphi)K_r(0, 0)\\Delta u\\Delta\\varphi, \\label{gamr} \\end{eqnarray} where $S_r$ and $K_r$ are defined as \\begin{equation} S_r (u, \\varphi) = \\Sigma (u, \\varphi) ~e^{u/2}~~~~\\rm and~~~~K_r (u, \\varphi) = \\frac{1+B^2 -e^{-u}\\cos(\\varphi)}{\\{ 2(\\cosh(u) -\\cos(\\varphi)) + B^2 e^u \\}^{3/2}}. \\label{SrKr} \\end{equation} In Eqs.~(\\ref{gamr}) and~(\\ref{SrKr}), $G$ denotes the gravitational constant, $u_{\\rm max} = \\log(r_{\\rm max}/r_{\\rm min})$ with $r_{\\rm max}$ the outer edge radius of the grid, $\\Sigma$ is the disk surface density, $\\Delta u$ and $\\Delta\\varphi$ are the mesh sizes, $K_r (0, 0)=1/B$ and $B=\\varepsilon_{\\rm sg} / r$. Since $\\varepsilon_{\\rm sg} \\propto r$ (see section~\\ref{sec:implement}), $B$ is uniform over the grid. The second term on the R.H.S. of Eq.~(\\ref{gamr}) is an additional corrective term that ensures the absence of radial self-force. Similarly, $g_{\\varphi} (u, \\varphi)$ reads \\begin{equation} g_{\\varphi} (u, \\varphi) = -G e^{-3u/2} ~\\int_{0}^{u_{\\rm max}} \\int_0^{2\\pi} S_{\\varphi}(u^{'}, \\varphi^{'})~K_{\\varphi}(u-u^{'}, \\varphi-\\varphi^{'}) ~du^{'} d\\varphi^{'}, \\label{gamt} \\end{equation} with $S_{\\varphi}$ and $K_{\\varphi}$ given by \\begin{equation} S_{\\varphi} (u, \\varphi) = \\Sigma (u, \\varphi) ~e^{3u/2}~~~~\\rm and~~~~K_{\\varphi} (u, \\varphi)=\\frac{\\sin(\\varphi)}{ \\{ 2(\\cosh(u) - \\cos(\\varphi)) + B^2 e^u \\}^{3/2} }. \\label{StKt} \\end{equation} In the particular case where only the axisymmetric component of the disk self-gravity is accounted for, which involves the axisymmetric component of the disk surface density $\\overline{\\Sigma}(u) = (2\\pi)^{-1} \\int_0^{2\\pi} \\Sigma(u,\\varphi)d\\varphi$, $g_{\\varphi}$ cancels out and \\begin{equation} g_r (u) = -G e^{-u/2}\\,\\int_{0}^{u_{\\rm max}} \\overline{S_r}(u^{'})\\,\\widetilde{K_r}(u-u^{'})\\,du^{'} + G \\overline{\\Sigma}(u) \\Delta u \\widetilde{K_r}(0), \\label{gamrzm} \\end{equation} where $\\overline{S_r}(u) = (2\\pi)^{-1} \\int_0^{2\\pi} S_r (u,\\varphi)d\\varphi$ and $\\widetilde{K_r}(u) = \\int_0^{2\\pi} K_r (u,\\varphi)d\\varphi$. ", "conclusions": "" }, "0801/0801.4139_arXiv.txt": { "abstract": "{Determining the evolutionary stage of a Young Stellar Object (YSO) is of fundamental importance to test star formation theories. Classification schemes for YSOs are based on evaluating the degree of dissipation of the surrounding envelope, whose main effects are the extinction of the optical radiation from the central YSO and re-emission in the far--infrared to millimeter part of the electromagnetic spectrum. Since extinction is a property of column density along the line of sight, the presence of a protoplanetary disk may lead to a misclassification of pre--main sequence stars with disks when viewed edge--on.} {We performed radiative transfer calculations to show the effects of different geometries on the main indicators of YSO evolutionary stage. In particular we tested not only the effects on the infrared colors, like the slope \\alf \\ of the flux between 2.2 and 24~\\mum , but also on other popular indicators of YSO evolutionary stage, such as the bolometric temperature and the optical depth of silicates and ices.} {We used the axisymmetric 3D radiative transfer codes RADMC and RADICAL to calculate the spectral energy distribution including silicates and ice features in a grid of models covering the range of physical properties typical of embedded and pre--main sequence sources.} {Our set of models compares well with existing observations, supporting the assumed density parametrization and the adopted dust opacities. We show that for systems viewed at intermediate angles (25\\degr\\ - 70\\degr ) the ``classical'' indicators of evolution are able to classify the degree of evolution of young stellar objects since they accurately trace the envelope column density, and they all agree with each other. On the other hand, edge-on system are misclassified for inclinations larger than $\\sim 65$\\degr $\\pm 5$\\degr\\ , where the spread is mostly due to the range of mass and the flaring degree of the disk. In particular, silicate emission, typical of pre--main sequence stars with disks, turns into silicate absorption when the disk column density along the line of sight reaches $1 \\cdot 10^{22}$\\persc, corresponding e.g. to a $5 \\cdot 10^{-3}$ \\msun \\ flaring disk viewed at 64\\degr. A similar effect is noticed in all the other classification indicators studied: \\alf , \\tbol , and the \\h2o\\ and \\co2\\ ice absorption strengths. This misclassification has a large impact on the nature of the flat--spectrum sources ($\\alpha \\simeq 0$), whose number can be explained by simple geometrical arguments without invoking evolution. A reliable classification scheme using a minimal number of observations is constituted by observations of the millimeter flux with both a single dish and an interferometer.} {} ", "introduction": "A correct classification of the evolutionary stages of young stellar objects (YSOs) is fundamental to understand the time--scales involved in the star formation process. Since protostars gradually accrete and disperse their envelope material during their lifetime, the amount of residual envelope material can be used to measure the degree of evolution of the protostellar object. The presence of a surrounding envelope has several characteristic effects on the observables of the YSO which can then be linked to the evolutionary stage of the system. In particular, the envelope strongly affects the spectral energy distribution (SED) of the protostellar system, extinguishing and reprocessing the protostellar radiation. Thus, instead of a single blackbody emission, the SED of a young protostellar system appears as a combination of two thermal components: one corresponding to the envelope emission ($T_{\\rm eff}< 100$~K) and the other that is the result of the emission of the central protostellar object ($T_{\\rm eff}> 2000$~K) extinguished and scattered by the surrounding envelope. It is evident that measuring the ``relative intensity'' of the two components would furnish the contribution of the envelope to the whole system. Indeed, low--mass protostellar systems were originally classified according to the amount of extinction towards the protostar, using the ratio between near (2~\\mum ) and mid-infrared fluxes (25-60~\\mum ) \\citep[\\alf$=d{\\mathrm log}(\\lambda F_{\\lambda})/d{\\mathrm log}(\\lambda)$,][]{adams1987,greene1994}, or by the relative amount of emission from the envelope, evaluated by the average frequency of emission ($\\left< \\nu \\right>=\\int \\nu F_{\\nu} d\\nu /\\int F_{\\nu} d\\nu$) converted to a blackbody temperature \\citep[\\tbol,][]{myers1993}. The combination of observations and theory has led to a three-class classification with class I representing the least evolved protostars with a significant fraction of the mass still left in the envelope and showing \\alf$>$0 and \\tbol$<$650 K; class II, corresponding to the classical T Tauri stars with disks dominating the infrared emission and with $-2<$\\alf$<0$ and 650~K$<$\\tbol$<$2800~K; and class III pre-main sequence stars which show the presence of circumstellar disks only in the mid--infrared and have \\alf$<-2$ and \\tbol$>$2800 K. The earlier class 0 phase was added by \\citet{andre1993} to include the most deeply embedded protostars that are invisible in the infrared. Finally, some authors refer to the sources with $-0.3<$\\alf$<0.3$ as flat--spectrum sources and view them as transitional objects between class I and II \\citep[e.g.,][]{greene1994}. This classification has had an enormous success and a huge impact on star formation studies, but it is subject to misclassification. In particular, it does not take into account that the dissipation of the envelope is not a spherical process; in fact, while part of the material is dispersed in the outflow, the bulk of it is accreted into a disk, with the result that systems viewed edge--on would present extinction values typical of younger systems with more massive envelopes, thus leading to a misclassification. This effect is clearly shown in the theoretical work of, e.g., \\citet{whitney2003a, whitney2003b} and \\citet{robitaille2006}. To emphasize the difference between the observational classification and the intrinsic properties of the YSO, \\citet{robitaille2006} introduced the alternative nomenclature ``stage I, II, III\", equivalent to the Lada classes, but referring to the true nature of the source regardless of its observed properties. These authors distinguish stage I from stage II YSOs according to their accretion rate of the envelope, with the dividing line set at \\mdot $_{\\rm env}=10^{-6}$~\\msun ~$\\rm yr^{-1}$ (stage II and III are defined according to the accretion rate of the disk instead of that of the envelope). In the equations for a rotationally flattened envelope, it easily can be seen that \\mdot $_{\\rm env}$ is ultimately a measure of the volume density of the envelope at the centrifugal radius; for a 1~\\msun \\ star with a 300~AU centrifugal radius \\mdot $_{\\rm env}=10^{-6}$~\\msun ~$\\rm yr^{-1}$ is equivalent to an H$_2$ number density of $1.8 \\cdot 10^5$~\\percc . \\\\ A possibility to observationally disentangle the extinction coming from envelopes to that produced in the disks can come from their very different temperature structures, with disks going from 30 to 1000~K and envelopes from 10 to 200~K. With such different conditions, dust grains are expected to be very different. Specifically, the envelope grains are expected to be coated by ice unlike the fraction of disk grains with temperatures above 100~K. Thus, comparing the evolutionary stage inferred by the infrared colors, like \\alf , with \\tbol\\, or the silicates and ice optical depths could provide a tool to distinguish between stage I YSO and edge--on disks. Including the presence of ices (\\h2o\\ and \\co2) in our opacity tables enables us to predict not only the infrared colors, as efficiently presented in the recent literature \\citep{robitaille2006}, but also the optical depth of spectral features, in an attempt to broaden the number of observables for YSO evolution diagnostics. In this paper, we present model predictions for a series of YSOs covering a large range of disk mass, envelope mass and stellar luminosity, and exploring the effect of inclination on the emerging SED and spectral features. Section~\\ref{mod} presents the radiative transfer model we used and the adopted ice--coated dust properties; Section~\\ref{res} compares the results of the calculation with existing data; Section~\\ref{con} summarizes our conclusions. ", "conclusions": "\\label{con} We used a Monte Carlo-based radiative transfer code to calculate the emission for a grid of models representing stage I and II young stellar objects: embedded YSOs and pre--main sequence stars with disks. The aim was to study the behaviour of classification parameters that are used to distinguish YSOs embedded in envelopes from those surrounded by disks only. We can summarize our results in these main points: \\begin{itemize} \\item The classical indicators of YSO evolution, \\alf ,\\tbol , and the silicate emission/absorption can describe very well the disappearance of the protostellar envelope for those systems seen at intermediate inclinations (approximately from 25\\degr\\ to 65\\degr ), accounting for more than half of the sources. Interestingly the transition between class I and class II measured with \\alf , \\tbol , or with the silicate feature happens for the same line of sight column density of $\\sim1 \\cdot 10^{22}$~\\persc \\ (corresponding to \\menv $\\sim$ 0.2~\\msun ), which means that these methods are well calibrated relative to each other. \\item Each of these indicators confuses pre--main sequence stars with edge-on disks with heavily embedded YSOs if the column density exceeds $1 \\cdot 10^{22}$~\\persc ; this column density is reached for inclinations greater than 60\\degr --70\\degr , depending on the mass of the disk. \\item Even a pre--main sequence star surrounded by disk of only $5 \\cdot 10^{-4}$ \\msun\\ is classified as a flat--spectrum source for inclinations greater than 70\\degr. This means that 34\\% of the stage II sources are misclassified and could account for the entire population of flat--spectrum sources observed in nearby star--forming regions. If 10\\% of pre--main sequence disk objects are misclassified as class I, the derived time-scale for the embedded phase is overestimated by a factor of two, when using the ratio of class I/II blindly. \\item \\co2\\ and \\h2o\\ ice optical depth correlate well with \\alf\\ and \\tbol, providing an alternative way to classify YSOs. For edge-on sources, the strong effect of scattering saturates the continuum around the features, thus affecting the feature strength and hiding the effect of temperature structure along the line of sight. \\item A combination of single dish and interferometer millimeter continuum observations is able to trace accurately, and with the minimum number of observations, the amount of remaining envelope in emission and thus the true evolutionary stage of the YSO. \\end{itemize}" }, "0801/0801.1236_arXiv.txt": { "abstract": "{The blazar \\object{AO 0235+164} was claimed to show a quasi-periodic behaviour in the radio and optical bands in the past, with the main outbursts repeating every 5--6 years. However, the predicted 2004 outburst did not occur, and further analysis suggested a longer time scale, according to which the next event would have occurred in the 2006--2007 observing season. Moreover, an extra emission component contributing to the UV and soft X-ray flux was detected, whose nature is not yet clear. An optical outburst was observed in late 2006 -- early 2007, which triggered a Whole Earth Blazar Telescope (WEBT) campaign as well as target of opportunity (ToO) observations by the Swift satellite.} {In this paper, we present the radio-to-optical data taken by the WEBT together with the UV data acquired by the UltraViolet and Optical Telescope (UVOT) instrument onboard Swift to investigate both the outburst behaviour at different wavelengths and the nature of the extra emission component.} {Multifrequency light curves have been assembled with data from 27 observatories; optical and UV fluxes have been cleaned from the contamination of the southern active galactic nucleus (AGN). We have analysed spectral energy distributions at different epochs, corresponding to different brightness states; extra absorption by the foreground galaxy has been taken into account.} {We found the optical outburst to be as strong as the big outbursts of the past: starting from late September 2006, a brightness increase of $\\sim 5$ mag led to the outburst peak in February 19--21, 2007. We also observed an outburst at mm and then at cm wavelengths, with an increasing time delay going toward lower frequencies during the rising phase. Cross-correlation analysis indicates that the 1 mm and 37 GHz flux variations lagged behind the $R$-band ones by about 3 weeks and 2 months, respectively. These short time delays suggest that the corresponding jet emitting regions are only slightly separated and/or misaligned. In contrast, during the outburst decreasing phase the flux faded contemporaneously at all cm wavelengths. This abrupt change in the emission behaviour may suggest the presence of some ``shutdown\" mechanism of intrinsic or geometric nature. The behaviour of the UV flux closely follows the optical and near-IR one. By separating the synchrotron and extra component contributions to the UV flux, we found that they correlate, which suggests that the two emissions have a common origin.} {} ", "introduction": "\\object{AO 0235+164} at redshift $z=0.94$ is one of the best-studied BL Lac objects. The analysis of its radio and optical light curves extending over 25 years led \\citet{rai01} to suggest a quasi-periodic occurrence of the main outbursts every $5.7 \\pm 0.5$ years. The next outburst was predicted around February--March 2004, and a multiwavelength campaign was organised by the Whole Earth Blazar Telescope (WEBT)\\footnote{{\\tt http://www.to.astro.it/blazars/webt/} \\\\ \\citep[see e.g.][]{vil06,vil07,boe07,rai07b}} to follow the expected event. The radio-to-optical observations by the WEBT were complemented by the optical--UV and X-ray data acquired by the XMM-Newton satellite during 3 pointings in January and August 2004, and January 2005, and by optical spectroscopic observations with the 3.6 m Telescopio Nazionale Galileo (TNG). The results of this intense observing effort were published by (\\citealt{rai05,rai06b,rai06a,rai07a}; see also \\citealt{hag07b} for an analysis of colour variability). The predicted outburst was not observed, and time-series analysis on the light curves extended to 2005 revealed a possible characteristic variability time scale of $\\sim 8$ years. Moreover, the XMM-Newton observations suggested the presence of an extra emission component in the source spectral energy distribution (SED), in addition to the synchrotron and inverse-Compton ones. The origin of this component, peaking in the UV/soft X-ray frequency range, was ascribed either to thermal emission from an accretion disc, or to synchrotron emission from an inner jet region. An increased radio activity was detected in the 2005--2006 observing season \\citep{bac07}, and a dramatic optical brightening was finally observed in late 2006 -- early 2007. This triggered a new WEBT multiwavelength campaign as well as ToO observations by the Swift satellite. In this paper, we present the radio-to-optical observations performed from spring 2005 (i.e.\\ the end of the period considered in \\citealt{rai06b}) to October 2007, and the data acquired by the UltraViolet and Optical Telescope (UVOT) instrument onboard Swift. X-ray data from the Swift X-ray Telescope (XRT) and Burst Alert Telescope (BAT) instruments will be presented in another paper (Kadler et al., in preparation). This paper is organised as follows: the procedures we adopted to treat the WEBT and UVOT data are outlined in Sect.~2, with particular attention paid to the subtraction of the southern AGN contribution from the optical and UV photometry. The multifrequency light curves are presented and discussed in Sect.~3 while in Sect.~4 time lags among variations at different frequencies are derived by means of statistical analysis. Spectral energy distributions with simultaneous data from near-IR to UV are constructed in Sect.~5, with the aim of separating the synchrotron and the extra emission components and of understanding their relationship. Finally, the results of our work are discussed in Sect.~6. ", "conclusions": "The claim by \\citet{rai01} of a possible quasi-periodic occurrence of the major radio and optical outbursts of AO 0235+164 every $5.7 \\pm 0.5$ years led to a mobilization of the international blazar community to observe the next event, predicted to peak in early 2004. A WEBT campaign was organised, with a huge international participation \\citep{rai05,rai06b,rai06a,rai07a}, but the source remained in a faint state and time series analysis on the updated light curves suggested a possible longer period of about 8--8.5 years, delaying the occurrence of the next outburst toward the 2006--2007 observing season. An increased radio activity was registered in 2005--2006 \\citep{bac07}, and a major optical outburst was finally observed in late 2006 -- early 2007, about 8.5 years after the previous major outburst peak, thus confirming the \\citet{rai06b} prediction. We observed the same event simultaneously in the near-IR and UV frequency ranges, and then at millimetric and centimetric wavelengths, with a progressive time delay toward lower frequencies. We estimated that the 1 mm and 37 GHz outbursts lagged behind the $R$-band one by about 3 weeks and 2 months, respectively. This latter lag appears a factor $\\sim 2$ longer than previously estimated \\citep{rai05,rai06b} when considering the historical light curves until spring 2005. This may reflect some real change either in the jet structure at still unresolved scales or in its energetics with respect to the past. Indeed, there are sources that exhibit a characteristic behavior for many years and then suddenly change \\citep[see e.g.][]{smi96,all96}. Another possibility is that lack of data for long time intervals, chiefly in the optical bands because of solar conjunctions, affects the results of the DCF analysis. Even in the case of the 1997--1998 outburst, which was intensively observed thanks to the efforts of the just-born WEBT \\citep{rai01}, the central part of the event was missed in the optical, and the dimming phase was poorly sampled at 37 GHz. As for the 2006--2007 outburst, notwithstanding the exceptional sampling, we cannot rule out that solar conjunction hid a further optical peak. This would shorten the radio lags. In any case, the delays estimated in the present paper appear rather short if compared, for example, with the approximately twice-longer lags estimated by \\citet{vil07} for the quasar-type blazar 3C 454.3 in correspondence of its 2004--2006 exceptional outburst, and by \\citet{vil04b} and \\citet{bac06} for BL Lacertae. The short time delays in AO 0235+164 had already been noticed in the past \\citep{web00,rai01}, and this was one of the reasons why microlensing was proposed as a possible explanation of the major variability events in this source \\citep[see e.g.][and references therein]{rai07a}. In contrast, when interpreting the multifrequency variability of AO 0235+164 in the framework of the helical jet model by \\citet{vil99}, the short time delays imply that the corresponding jet-emitting regions are only slightly misaligned \\citep{ost04}. If, besides the geometrical effect, the outburst is also produced by energetic processes inside the jet, short time lags imply that the emitting regions are also closeby. At lower ($< 37$ GHz) radio frequencies, the outburst behaviour changes: while the rising phase is progressively delayed, as expected, the decaying phase is observed simultaneously at all wavelengths. In particular, the 22 GHz flux density reaches a maximum value similar to the 37 GHz flux density almost at the same time and then suddenly decreases. Although the observed general multifrequency properties of the outburst can be explained by models dealing with shocks propagating along an inhomogeneous jet \\citep[see e.g.][]{hug89,val92}, the emission behaviour during the dimming phase might suggest that the scenario is more complicated than what is depicted in these models. Indeed, it seems as if a kind of shutdown occurred in the jet between the mm emission region and the 37--22 GHz region. We can speculate that either a jet bending or an intrinsic power off of the perturbation may produce this effect, which could also account for the different flavours of events observed in the historical light curves of the source: the harder outbursts, which are stronger at the higher frequencies, suggest that a shutdown (of any origin) has occurred between the higher- and the lower-frequency emitting regions. Understanding whether the geometric or intrinsic scenario would be more plausible is not easy. One can envisage that if the flux fading is due to a jet bending, the perturbation could continue to travel along the jet, following a misaligned path, until it could eventually be observed at radio wavelengths as soon as the jet (helical) path turns again toward the line of sight. For those sources, for which the Very Long Baseline Interferometry (VLBI) resolution can separate these regions, we should be able to observe a brightening of a VLBI knot. We notice that our analysis of the multifrequency light curves in terms of harder/softer events presents some similarities with the generalized shock model by \\citet{val92}, who distinguish between high- and low-peaking flares. We finally mention that \\citet{hag07a} interpreted the photometric and polarimetric variations of AO 0235+164 during the December 2006 flare as due to the propagation of a transverse shock, accompanied by a small change in the viewing angle of the jet. Another major issue we investigated is the existence of an extra emission component mostly contributing in the UV and soft X-ray energy range of the source SED. Swift-UVOT observations during the culminating phase of the optical outburst show that the behaviour of the UV emission strictly follows that of the optical one. Moreover, when constructing SEDs with simultaneous near-IR-to-UV data, and separating the synchrotron and extra-component contributions, we found that they are correlated. Although this result is affected by uncertainties due to the decomposition procedure, it nevertheless suggests that an interpretation of the extra component in terms of radiation from the accretion disc is rather unlikely \\citep[see][for the various possible interpretations]{rai06b,rai06a}. Also the hypothesis of two independent synchrotron components appears now inadequate. A still viable explanation could be that of anomalous absorption of the optical to near-UV emission in a spectrum that would otherwise be power-law-like from the near-IR to UV. Indeed, examples of noticeable intrinsic absorption are found when analysing the spectra of broad absorption line (BAL) quasars, where several lines between $\\sim$ 1000 and 1500 \\AA\\ (rest frame) can heavily reduce the flux in the corresponding spectral regions \\citep[see e.g.][]{tur88}. A further possibility is that the extra component is the result of inverse-Compton scattering of radio photons off the relativistic electrons producing the IR--optical emission. Indeed, as stated above, the close correlation and short time delays between the optical and radio variations suggest that a lot of radio photons are available in the optical emitting region." }, "0801/0801.3469_arXiv.txt": { "abstract": "We introduce a differential equation for star formation in galaxies that incorporates negative feedback with a delay. When the feedback is instantaneous, solutions approach a self-limiting equilibrium state. When there is a delay, even though the feedback is negative, the solutions can exhibit cyclic and episodic solutions. We find that periodic or episodic star formation only occurs when two conditions are satisfied. Firstly the delay timescale must exceed a cloud consumption timescale. Secondly the feedback must be strong. This statement is quantitatively equivalent to requiring that the timescale to approach equilibrium be greater than approximately twice the cloud consumption timescale. The period of oscillations predicted is approximately 4 times the delay timescale. The amplitude of the oscillations increases with both feedback strength and delay time. We discuss applications of the delay differential equation (DDE) model to star formation in galaxies using the cloud density as a variable. The DDE model is most applicable to systems that recycle gas and only slowly remove gas from the system. We propose likely delay mechanisms based on the requirement that the delay time is related to the observationally estimated time between episodic events. The proposed delay timescale accounting for episodic star formation in galaxy centers on periods similar to $P\\sim 10$ Myrs, irregular galaxies with $P\\sim 100$ Myrs, and the Milky Way disk with $P\\sim 2$ Gyr, could be that for exciting turbulence following creation of massive stars, that for gas pushed into the halo to return and interact with the disk and that for spiral density wave evolution, respectively. ", "introduction": "Gas present in a galaxy fuels star formation or nuclear black hole growth. However both star formation and active galactic nuclei then release energy and momentum into the interstellar medium (ISM). Consequently the activity can suppress subsequent star formation. The process in which part of the output of a system is returned to its input and influences its further output is termed ``feedback.\" Early studies showed that when feedback by radiative heating is taken into account during gas accretion onto a central mass, steady solutions may not exist \\citep{ostriker76} and the feedback process can cause oscillations or periodic bursts of accretion \\citep{cowie78}. Simulations taking into account feedback processes illustrate that gas flows and star formation in galaxies can exhibit episodic or cyclic behavior \\citep{dong03,pelupessy04,ciotti07,stinson07} or alternatively can asymptotically converge onto a self-regulated equilibrium state \\citep{andersen00,robertson08}. Galaxies display complex star formation histories. Studies of irregular galaxy populations (e.g., \\citealt{tosi91,smeckerhane94,dohmpalmer02,dolphin03,skillman05,young07,dellenbusch08}), the Milky Way disk \\citep{rochapinto00}, galaxy centers \\citep{bland03,walcher06,cecil01} and the statistics of distant galaxies \\citep{glazebrook99} infer that multiple events of vigorous star formation, separated by millions to billions of years, can occur even in isolated galactic systems. Other studies (e.g., \\citealt{vanzee01,skillman05}) find little evidence for episodic star formation. However, theoretical work has primarily focused on self-regulated star formation \\citep{andersen00,silk01,elmegreen02,monaco04,krumholz06,slyz05,li06,dib06,joung06,elmegreen07,booth07,wada07,schaye07,robertson08} and has not explored when episodic rather than a steady rate of star formation is expected. As gas flows involving energy input, heating and cooling are complex, there is no simple way to predict when behavior is episodic or cyclic. However it is possible that average quantities can be estimated for these flows and relations based on these quantities can be used to classify their behavior. Delay differential equations can exhibit solutions that asymptotically approach a self-limiting equilibrium state and those that are periodic, even when feedback is negative. Consequently these equations can be used to differentiate between these two behaviors. Delay differential equations have been used to model biological systems with delayed negative feedback (e.g., \\citealt{wazewska88,gyori91,gurney80,kulenovic89}) but have not been applied to astrophysical systems. In this paper, using a delay differential equation, we determine when cyclic or periodic behavior is exhibited by the solutions rather than a smooth decay to a self-regulated steady state. We apply the theory to star forming galaxies, identifying delay mechanisms that could account for episodic accretion events inferred from observations. ", "conclusions": "It is now widely recognized that a detailed understanding of feedback and accretion processes is essential to progress in many fields of astrophysics and across the entire cosmological hierarchy, from galaxy clusters down to the scales of individual star forming regions. In order to progress, we will need huge improvements in analytic algorithms and computer power, as well as better conceptual tools for classifying complex behavior. Some processes may indeed be episodic or cyclic, while other instances may exhibit quasi-periodic cycles on the way to fully chaotic behavior. A deeper understanding requires that we should to some degree be able to distinguish between these two very different dynamical manifestations for open and closed systems. Here we have introduced a simple differential equation model that captures some of the complexity exhibited by astrophysical star forming systems with feedback. We introduce a one dimensional DDE for the molecular cloud density that allows cloud formation to depend on the star formation rate but at a previous time. Thus current star formation only affects the cloud distribution at a future time, we denote the delay time. The feedback is negative, so in the absence of delay there are no cyclic solutions or instabilities and all solutions asymptotically approach a self-limiting value. We illustrate that even when the feedback is negative a delay can cause cyclic or episodic behavior. The DDE captures phenomena exhibited by astrophysical simulations of this process, including periodic solutions in some cases but not in others. The DDE allows us for the first time to classify the solutions and predict when an astrophysical system is self-limiting or likely to exhibit periodic behavior based on timescales that are related to physical feedback and star formation processes. We find that periodic behavior is likely when two conditions are met. First, the delay timescale must exceed the cloud consumption timescale. Secondly, the star formation must be effective at reducing the rate of formation at densities near the self-limiting or steady state value. This is equivalent to requiring strong feedback or to requiring that the timescale to approach equilibrium be larger than approximately twice the cloud consumption timescale. We find that the amplitude of the oscillations is sensitive to the feedback strength and to a lesser extent on the ratio of the delay time to the consumption timescale. We focus on the molecular or self-gravitating cloud density in a galaxy as the most likely variable for the DDE. This allows recycling of gas over long periods of time as gas is recycled through clouds much faster than it is depleted by star formation. The consumption timescale is set by the lifetime of molecular clouds. When feedback delay times are longer than this timescale we predict episodic star formation events and with a period approximately 4 times the delay timescale. At the present time, there are no compelling constraints on either the feedback strength or the delay time, i.e. the two key parameters of the DDE model. Thus, it is difficult to apply the model rigorously although we suggest avenues for further exploration. There is more than one candidate for the delay time and associated feedback mechanisms, in particular, the timescale for supernovae to contribute to turbulence, the timescale for spiral density waves to evolve, and the timescale for material sent into the halo to return to interact with the disk. We associate these three candidate delay mechanisms with possible explanations for episodic star formation events in galaxy centers (on 10 Myr timescales), the solar neighborhood (on Gyr timescales) and dwarf galaxies (on 100 Myr timescales), respectively. Using a log normal density distribution we estimate that feedback is likely to be strong enough that the second condition for episodic solutions can be satisfied. The approach outlined here is potentially powerful framework to interpret and motivate future observations and simulations. Similar models might be applied to other accreting systems with feedback such as cooling flows. With better observationally constrained models we may be able to use similar simple dynamical models as recipes to drive simulations or interpret statistics of astrophysical objects that exhibit episodic accretion. Lacking currently are simulations and observational programs that constrain the timescales and strengths of possible feedback mechanisms and their functional form. In view of this uncertainty, we adopted an exponential function for the feedback process, but is this fully justified? Evidence for feedback-influenced star formation could be sought by probing for correlations between turbulence and deviations from empirical star formation laws. Other forms for the feedback function could be used, such as that of the Mackey-Glass model which can exhibit chaotic behavior \\citep{glass}. More sophisticated global theories of star formation could be developed to better predict the form of the feedback and go beyond the self-limiting equilibrium state models. Higher dimensional models could be explored, similar to those used to model predator and prey populations. By going to systems with additional variables it should be possible to model these systems without delays. The period is not strongly dependent on the amplitude of oscillation for the simple model explored here, however, this may not be true for more complex models. \\vskip 0.5 truein We thank Adam Frank, Eric Blackman, Jason Nordhaus and Richard Edgar for helpful discussions. Support for this work was in part provided by by NASA through awards issued by JPL/Caltech, National Science Foundation grants AST-0406823 $\\&$ PHY-0552695, the National Aeronautics and Space Administration under Grant No.$\\sim$NNG04GM12G issued through the Origins of Solar Systems Program, and HST-AR-10972 to the Space Telescope Science Institute. JBH is funded by a Federation Fellowship from the Australian Research Council. \\appendix" }, "0801/0801.4763_arXiv.txt": { "abstract": "We investigate the apparent discrepancy between gas and dust outer radii derived from millimeter observations of protoplanetary disks. Using 230 and 345~GHz continuum and CO J=3-2 data from the Submillimeter Array for four nearby disk systems (HD 163296, TW Hydrae, GM Aurigae, and MWC 480), we examine models of circumstellar disk structure and the effects of their treatment of the outer disk edge. We show that for these disks, models described by power laws in surface density and temperature that are truncated at an outer radius are incapable of reproducing both the gas and dust emission simultaneously: the outer radius derived from the dust continuum emission is always significantly smaller than the extent of the molecular gas disk traced by CO emission. However, a simple model motivated by similarity solutions of the time evolution of accretion disks that includes a tapered exponential edge in the surface density distribution (and the same number of free parameters) does much better at reproducing both the gas and dust emission. While this analysis does not rule out the disparate radii implied by the truncated power-law models, a realistic alternative disk model, grounded in the physics of accretion, provides a consistent picture for the extent of both the gas and dust. ", "introduction": "Characterizing the gas and dust distribution in the disks around young stars is important for understanding the planet formation process, as these disks provide the reservoirs of raw material for nascent planetary systems. A common method of modeling circumstellar disk structure is to use models described by power laws in surface density and temperature that are truncated at a particular outer radius. This prescription has its historical roots in calculations of the minimum mass solar nebula, which indicated a surface density profile of $\\Sigma \\propto r^{-3/2}$ \\citep[e.g.][]{wei77}, as well as theoretical predictions of a radial power-law dependence of temperature for accreting disks around young stars \\citep{ada86,ada87}. Observationally, the parameterization of temperature and surface density as power-law functions of radius began with early spatially unresolved studies of continuum emission from disks \\citep{bec90,bec91}. These models have since been refined and applied to spatially resolved observations of many disks with success \\citep[e.g.][]{mun93,dut94,lay94,dut98}, and they have proven useful for understanding the basic global properties of disk structure. Recently, however, with the advent of high signal-to-noise, multi-frequency observations of gas and dust in protoplanetary disks, these models have begun to encounter difficulties, particularly in the treatment of the outer disk edge. The extent of the gas and dust distribution in circumstellar disks has implications for our understanding of the planet formation process in our own solar system. There is some evidence for a sharp decrease in the surface density of Kuiper Belt objects beyond a distance of 50 AU from the Sun \\citep{jew98,tru01,pet06}. However, the origin of this edge is unclear. \\citet{ada04} note that the observed distance is far interior to the radius at which truncation by photoevaporation would be expected to occur, while \\citet{you02} find that the presence of such an edge in planetesimal density could be explained by drift-induced enhancement. A compelling possibility is that the Sun formed in a cluster environment, and the early solar disk was truncated by a close encounter with a passing star \\citep[see][and references therein]{rei05}. A more complete understanding of the outer regions of protoplanetary disks may provide insight into the processes that shape the outer solar system. \\citet{pie05} present multiwavelength millimeter continuum and CO isotopologue observations of the disk around the Herbig Ae star AB Aurigae and found from fitting models of disk structure described by truncated power laws that the outer radius of the dust derived from continuum emission ($350\\pm30$~AU) was much smaller than that of the gas derived from $^{12}$CO J=2-1 emission ($1050\\pm10$~AU). They suggest that a change in dust grain properties resulting in a drop in opacity could be responsible for the difference, and note the possible association with a ring feature in the disk at 200~AU. A similar result was obtained by \\citet{ise07} from observations of the disk around the Herbig Ae star HD 163296: they found a significant discrepancy between the outer radius derived for the dust continuum emission ($200 \\pm 15$ AU) and that derived from CO emission ($540 \\pm 40$ AU). These data appeared to require a sharp drop in surface density, opacity, or dust-to-gas ratio beyond 200~AU; however, as they discuss, there is no obvious physical basis for such a discontinuity. As \\citet{ise07} demonstrate, the discrepancy in outer radii derived from the dust and gas is not simply an issue of sensitivity; the observations were sufficiently sensitive to detect emission from the power-law dust disk if it did extend to the radius indicated by the CO emission. The underlying issue is that the truncated power law model does not simultaneously reproduce the extent of both the continuum and CO emission for these disks. Using data from the Submillimeter Array\\footnote{The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory and the Academia Sinica Institute of Astronomy and Astrophysics and is funded by the Smithsonian Institution and the Academia Sinica.} we show that the same apparent discrepancy in gas and dust outer radius applies to the circumstellar disks around several more young stars. In an attempt to understand the origin of this discrepancy, we investigate an alternative surface density profile based on work by \\citet{har98}, which is similar to a power law profile in the inner disk but includes a tapered outer edge. We show that this model, which has a physical basis in similarity solutions of disk evolution with time, is capable of simultaneously reproducing both continuum and CO emission from these disks. The primary difference between this model and the truncated power-law disk is that instead of a sharp outer edge the surface density falls off gradually, with sufficient column density at large radii that CO emission extends beyond the point at which dust continuum emission becomes negligible. ", "conclusions": "With the advent of high signal-to-noise interferometer observations that resolve the outer regions of nearby protoplanetary disks, an apparent discrepancy has emerged between the extent of the dust continuum and molecular gas emission \\citep{pie05,ise07}. Using multi-frequency interferometric data from the Submillimeter Array, we have investigated this disparity for four disk systems (HD 163296, TW Hydrae, GM Aurigae, and MWC 480) in the context of two distinct classes of disk structure models: (1) a truncated power law, and (2) a similarity solution for the time evolution of an accretion disk. The primary difference between these models is in their treatment of the disk outer edge: the abruptly truncated outer edge of the power-law disk causes the visibilities to drop rapidly to zero, leading to an inferred outer radius that is small in comparison with the observed molecular gas emission. The similarity solution, by contrast, tapers off smoothly, creating a broader visibility function and allowing molecular gas emission to persist at radii well beyond the region in the disk where continuum falls below the detection threshold. The outer radius discrepancy appears to exist only in the context of the power-law models. In light of this result, it appears that an abrupt change in dust properties for these disks is unlikely, as there is no physical mechanism to explain such a discontinuity. This may imply that a sharp change in dust properties in the early solar nebula is similarly an unlikely explanation for the Kuiper belt edge observed by \\citet{jew98}, and that a dynamical mechanism such as truncation by a close encounter with a cluster member \\citep[][and references therein]{rei05} may provide a more plausible origin. In this case, we would expect to observe disks with sharp outer edges only in clustered environments, and a model with a tapered edge would be a more realistic prescription for investigating the structure of a typical isolated disk. The tapered disk models provide a natural explanation for the disparate outer radii observed using different probes of the disk extent, including comparison of continuum and molecular gas observations \\citep{pie05,ise07}, and also comparison of different isotopologues and rotational transitions of a particular molecule \\citep{pie07}. When predicting CO emission, this simple model does neglect potential variance in the CO abundance due to depletion in the midplane and photodissociation at the disk surface; however, the results presented are intended simply to illustrate the global differences between gas and dust emission from the two model classes, independent of detailed CO chemistry. While we cannot rule out disparate gas and dust radii in these disks, we show that an alternative disk structure model, grounded in the physics of accretion, resolves the apparent size discrepancy without the need to invoke dramatic changes in dust opacity, dust density, or dust-to-gas ratio in the outer disk." }, "0801/0801.2765_arXiv.txt": { "abstract": "We report the discovery of quiescent emission from molecular hydrogen gas located in the circumstellar disks of six pre-main sequence stars, including two weak-line TTS, and one Herbig AeBe star, in the Chamaeleon~I star forming region. For two of these stars, we also place upper limits on the $2\\rightarrow1~S(1)$/$1\\rightarrow0~S(1)$ line ratios of $\\sim$~0.4 and 0.5. Of the 11 pre-main sequence sources now known to be sources of quiescent near-infrared hydrogen emission, four possess transitional disks, which suggests that detectable levels of H$_2$ emission and the presence of inner disk holes are correlated. These H$_2$ detections demonstrate that these inner holes are not completely devoid of gas, in agreement with the presence of observable accretion signatures for all four of these stars and the recent detections of [Ne~{\\scshape ii}] emission from three of them. The overlap in [Ne~{\\scshape ii}] and H$_2$ detections hints at a possible correlation between these two features and suggests a shared excitation mechanism of high energy photons. Our models, combined with the kinematic information from the H$_2$ lines, locate the bulk of the emitting gas at a few tens of AU from the stars. We also find a correlation between H$_2$ detections and those targets which possess the largest H$\\alpha$~equivalent widths, suggesting a link between accretion activity and quiescent H$_2$ emission. We conclude that quiescent H$_2$ emission from relatively hot gas within the disks of TTS is most likely related to on-going accretion activity, the production of UV photons and/or X-rays, and the evolutionary status of the dust grain populations in the inner disks. ", "introduction": "\\label{intro} The reservoirs of gas and dust found orbiting most sun-like stars with ages on the order of a million years --- also known as T Tauri stars (TTS) --- are now subjects of intense study, as they likely hold the secrets to the formation history of the solar system and other planetary systems. The evolution of the observable characteristics of these planet forming disks and the underlying phenomena responsible for their detectability may provide the necessary insight to further our understanding of planet forming processes and to determine the likelihood of the formation of planetary systems around sun-like stars. The existence of circumstellar disks accompanying young low mass stars previously has been inferred by the presence of detectable emission from trace constituents of these disks (i.e., dust and gaseous molecules other than H$_2$). The earliest attempts to determine the fraction of stars that possess disks in young star forming clusters focused on their near-infrared continuum \\citep{stro1989} and submillimeter continuum emission \\citep{wein1989,beck1990}. Circumstellar disks possessing warm ($T$~$\\sim$~1000~K) micron-sized dust grains within $\\sim$~1~AU of the pre-main sequence stars produce near-infrared emission in excess of that predicted for the stellar photosphere, while circumstellar disks possessing cold, micron-sized dust grains at radii of several tens of AU produce excess radiation at mid-infrared and submillimeter wavelengths. From surveys such as these, astronomers have determined the lifetime of warm dust grains in the inner disks to be 3-6~Myr \\citep{lada2000,hais2000,hais2001a}. Similarly, observations of cold ($T$~$\\sim$~10-100~K) gas molecules (e.g., CO, HCN$^+$, etc) from the outer regions of the disks \\citep{beck1993,zuck1995} tend to support the disk lifetime estimates drawn from near-infrared surveys. More recently, new disk fraction and lifetime surveys of TTS in nearby star forming regions have been conducted utilizing the Infrared Array Camera (IRAC) and Multiband Imaging Photometer (MIPS) aboard the {\\it Spitzer Space Telescope}. These instruments are capable of detecting emission from dust grains ($T$~$\\sim$~100-300~K) located at intermediate disk radii, a region of these disks to which the previous photometric surveys were insensitive. \\cite{ciez2007} took a census of 230 sources identified as weak-line TTS -- weakly or non-accreting TTS possessing either a passive, optically thin disk or no circumstellar material -- in the Lupus, Ophiuchus, and Perseus molecular clouds. They find a higher disk frequency for the sources located in the highest extinction regions of the clouds, suggesting that the wTTS at the edges of the formation regions are slightly older and more evolved. Although they find $\\sim$~20\\% of these possess some circumstellar material, $\\sim$~50\\% of what appear to be the youngest stars in the sample show no evidence of disks. These results, which are consistent with other {\\it Spitzer} surveys of low mass star forming regions \\citep{furl2006,geer2006,kess2006,padg2006,damj2007}, lead them to conclude, based on the existence of detectable emission from cool dust grains, that disk lifetimes are less than ten million years, which roughly agrees with the conclusions about disk lifetimes drawn from near-infrared surveys. Given the chaotic environment near these young stars and the phenomena (i.e., accretion, photoevaporation, stellar and disk winds, close encounters with other cluster members) that conspire to destroy these disks, the lifetimes measured from these surveys can be explained by disk dissipation models including all or some of these processes \\citep{holl2000}. Despite the growing consensus that disk lifetimes are only a few million years that has been drawn from several different techniques for studying circumstellar disks, astronomers must remain cautious when inferring the presence or absence of disks based upon the detectability of disk tracers (i.e., dust and CO emission) because the majority of the disk mass is composed of molecular hydrogen (H$_2$). As these disks evolve according to the core accretion model of planet formation, the small micron-sized dust grains accumulate into larger bodies, which will change the wavelength dependence of the dust emissivities and the collective surface area of dust grains emitting at wavelengths in the near-infrared. Since a planetesimal must reach a mass of $\\simeq$~10 earth masses before it can gravitationally collect H$_2$ gas from the disk, the main component of the disk must persist, while planetesimals accrete into km-sized objects and disk tracers disappear, in order for gas giant planets to be the result of the core accretion scenario of planet formation. Therefore, in an accretionally-evolved system in which small dust grains have coagulated into larger planetesimals, an evolutionary stage should exist during which both the near-infrared and mid-infrared excesses will have diminished substantially while the massive gaseous component remains present. In previous work, \\cite{wein2000}, \\cite{bary2002}, and \\cite{bary2003a} searched a variety of TTS, predominantly wTTS, in the nearby Taurus-Auriga, $\\rho$ Ophiuchus, and TW Hya Association star forming regions for emission signatures from H$_2$ persisting in such accretionally evolved circumstellar disks. Using various high-resolution, near-infrared spectrometers, narrow line emission centered at the $v=1\\rightarrow0~S(1)$ transition rest wavelength of 2.1218~$\\mu$m was detected from several TTS stars; some were known to possess circumstellar disks (TW~Hya, GG~TauA, and LkCa~15), and one was previously thought to be devoid of any circumstellar material (DoAr~21). The detection of quiescent emission from the disk of a wTTS with no other detectable disk tracers suggests that the gaseous component may indeed persist as the disk evolves. However, we estimate the mass of hot H$_2$ gas producing the observed features in these sources to be in the range 10$^{-12}$-10$^{-10}$~M$_\\odot$. For at least three of these four sources, a substantial reservoir of H$_2$ gas exists in their disks suggesting that the near-infrared H$_2$ emission feature, is itself only a tracer of the disk gas. In the time since these detections were made, several other authors searching for the same emission signature also have detected H$_2$ gas \\citep{itoh2003,taka2004,rams2007}. In two of the three additional sources toward which H$_2$ was detected (LkH$\\alpha$~264 and ECHA J0843.3-7905), the emission lines were similarly centered at 2.1218~$\\mu$m, with full-width at half of maximum (FWHM) only slightly larger than those reported for the previous four detections. In the case of DG~Tau, \\cite{taka2004} reported the detection of an H$_2$ emission line blueshifted by 15~km~s$^{-1}$ and possessing a FWHM of $\\sim$~35~km~s$^{-1}$. The authors also show that the emitting gas is extended along the axis of a strong outflow previously observed from this source. Therefore, the emitting gas from this source is most likely associated with the outflowing gas distinguishing this detection from the others. In addition to these detections of rovibrational emission, \\cite{sher2003}, \\cite{bitn2007} and \\cite{mart2007} have detected H$_2$ emission from the mid-infrared pure rotational transitions from the Herbig AeBe stars, AB~Aur and HD~97048, with a similarly quiescent kinematic signature. Here, we present five more detections of H$_2$ near-infrared emission from both classical and weak-line TTS, an infrared companion (IRC), and a detection from a Herbig AeBe star in the Chamaeleon I star forming region. In addition, we place an upper limit on emission from a second line of H$_2$ at 2.2477~$\\mu$m for two of these sources. We review the possible stimulation mechanisms with respect to recent X-ray and UV irradiated disk models. We find that two of our past detections and one of the present detections possess transitional disks --- disks with large inner disk holes devoid of emission from small warm dust grains. Two of these transitional disk sources also possess detectable emission from a mid-infrared [Ne~{\\scshape ii}] emission line predicted to be a tracer of high energy photons and a signature of photoevaporation, a mechanism thought to be potentially responsible for the presence of the inner disk holes of the transitional disks. Kinematic information from the H$_2$ emission features suggest that the emitting gas extends to within a few AU of each source, demonstrating for the sources with inner disk holes that gas remains despite the lack of dust emission. We estimate the location of the regions within the disks where potential excitation photons deposit their energies and find agreement with the location of emitting gas determined from the velocity dispersion of the line. Finally, we relate the detection of quiescent H$_2$ emission to accretion activity and, possibly, the evolutionary status of the dust grain population. ", "conclusions": "As part of a high resolution near-infrared spectroscopic survey for H$_2$ emission from TTS in nearby star forming regions, we have made new detections of emission features centered on the 2.1218~$\\mu$m $v=1\\rightarrow0~S(1)$ transition of H$_2$ in the spectra of two cTTS (CS~Cha and CV~Cha), two wTTS (Glass Ia and Sz~33), one IRC (Glass Ib), and one Herbig AeBe star (HD~97048) located in the Chamaeleon~I dark cloud. The detection of H$_2$ we report toward HD~97048 is the first detection of near-infrared emission from quiescent H$_2$ toward a Herbig AeBe star. In every case, the central velocity of the emitting gas is found to be coincident with the systemic velocity of the source. Assuming a circumstellar disk origin for these emission lines, an assumption well-supported by the central velocities of these lines falling at the rest velocities of the respective stars and by the narrow line widths, we used the HWZI and the FWHM velocities of these Gaussian-shaped emission features to estimate the innermost radii to which the gas extends and the radii at which the bulk of the emitting gas resides . We find that the innermost radii for all of the stars is on the order of $\\sim$~1~AU. We further find that the radii about which the bulk of the emission is centered is on the order of $\\sim$~10~AU. For the binary source (Glass I), H$_2$ emission extends between both components and to the south of the primary. Although the extended emission suggests that the gas could be entrained in an outflow, the lack of a detectable velocity gradient along the would-be axis of the outflow makes this an unlikely scenario. However, a circumbinary disk origin for the extended emission is also difficult to reconcile with the large FWHM of the emission lines measured for the extended emission. The initial goals for this survey was to determine if wTTS lacking detectable emission from standard disk tracers still harbor protoplanetary disks capable of or in the process of forming planetary systems. Our survey revealed two such stars, making a total of three, including DoAr~21 discovered by us in a previous survey. The range of H$_2$ line fluxes (6.9$\\times$10$^{-16}$~$\\le$~F$_{2.12}$~$\\le$~8.6$\\times$10$^{-15}$~ergs~cm$^{-2}$~s$^{-1}$) and masses (10$^{-10}$~$\\le$~M$_{H_2}$~$\\le$~10$^{-11}$~M$_\\odot$) measured for the Chamaeleon~I sources with detected H$_2$ line emission were found to agree with all previous detections of quiescent H$_2$ emission. These trace amounts of gas do not directly inform us about the total disk masses harbored by these systems, though almost certainly the mass of H$_2$ detected directly through line emission at 2.1218~$\\mu$m represents only a small fraction of the total mass in these disks. Our disk models suggest that the total disk masses could be 10$^2$ to 10$^7$ times greater than the mass measured directly in this single emission line. Assuming that shock excitation is not the most likely mechanism for stimulating narrow line emission from H$_2$ gas which is not Doppler shifted from the systemic velocity of the source, we suggest that UV fluorescence and/or X-rays are the most likely stimulation mechanisms responsible for the observed H$_2$ emission. We calculated optical depth surfaces for UV photons and X-rays penetrating a disk composed entirely of H$_2$ to determine where in a core-accretionally evolved disk these photons deposit their energy and possibly generate H$_2$ emission. These models confirm that the radiation is most likely to be absorbed by gas in the upper atmospheres of these disks at intermediate radii and show that these stimulation mechanisms require significantly dense and massive columns of gas to produce the observed emission features at radii coincident with the those determined from the kinematic information. Many improvements may be made to these calculations by including opacities associated with large dust grains, other gaseous species, or a disk model that better approximates an irradiated transitional disk. However, the results from these calculations suggest that a gas-rich inner disk is necessary to produce quiescent H$_2$ emission via UV photons and X-rays, and that the stimulated gas resides at high disk altitudes and intermediate disk radii. All four sources in our survey to-date known to possess transitional disks with optically thin inner disk holes are also H$_2$ detections. These detections place gas within the inner disk holes similar to the resolved and unresolved detections of [Ne {\\scshape ii}] emission lines from the sources. The presence of gas in the inner disk holes confirms the presence of gas within these gaps necessary to explain the accretion activity observed for these sources. The correlation between sources possessing quiescent H$_2$ emission and [Ne {\\scshape ii}] suggests a shared stimulation mechanism for both of these lines, a relationship to the photoevaporation process, and a distinct stage in the evolution of these circumstellar disks. We also find that most sources in our survey with H$\\alpha$~EWs sufficient to classify them as cTTS have detectable H$_2$ emission. The only exception is WW~Cha, which may also be a detection, though we have cautiously classified it as a `possible' detection, sine the H$_2$ emission towards this source, if real, is only at the 2.4$\\sigma$ level. The apparent relationship between H$\\alpha$ emission and accretion activity suggests that the quiescent H$_2$ emission may likewise be correlated with accretion in agreement with the findings of \\cite{carm2007}. This is not entirely surprising given that the UV photons necessary for stimulating H$_2$ emission are likely generated by accretion hot spots on the surfaces of the stars. In fact, this is strong evidence that UV fluorescence undoubtedly contributes to the observed H$_2$ emission. Given the dependence of the UV penetration depths on the sizes of the dust grains in the disk, illustrated by \\cite{nomu2007}, and the overlap between H$_2$ detections and sources possessing transitional disks, we suggest that the evolutionary stage of the dust grain population in the inner regions of these disks is also correlated with the production of quiescent H$_2$ emission. Based upon these findings, we propose that Sz~33 and DoAr~21, both H$_2$ sources that are borderline c/wTTS, likely harbor transitional disks." }, "0801/0801.3555_arXiv.txt": { "abstract": "{} { Observations of shell-type supernova remnants (SNRs) in the GeV to multi-TeV $\\gamma$-ray band, coupled with those at millimetre radio wavelengths, are motivated by the search for cosmic-ray accelerators in our Galaxy. The old-age mixed-morphology SNR W~28 (distance $\\sim$2~kpc) is a prime target due to its interaction with molecular clouds along its northeastern boundary and other clouds situated nearby.} { We observed the W~28 field (for $\\sim$40~h) at very high energy (VHE) $\\gamma$-ray energies ($E>$0.1~TeV) with the H.E.S.S. Cherenkov telescopes. A reanalysis of EGRET $E>$100~MeV data was also undertaken. Results from the NANTEN 4m telescope Galactic plane survey and other CO observations were used to study molecular clouds. } { We have discovered VHE $\\gamma$-ray emission (HESS~J1801$-$233) coincident with the northeastern boundary of W~28 and a complex of sources (HESS~J1800$-$240A, B and C) $\\sim$0.5$^\\circ$ south of W~28 in the Galactic disc. The EGRET source (GRO~J1801$-$2320) is centred on HESS~J1801$-$233 but may also be related to HESS~J1800$-$240 given the large EGRET point spread function. The VHE differential photon spectra are well fit by pure power laws with indices $\\Gamma \\sim 2.3$ to 2.7. The spectral indices of HESS~J1800$-$240A, B, and C are consistent within statistical errors. All VHE sources are $\\sim$10$^\\prime$ in intrinsic radius except for HESS~J1800$-$240C, which appears pointlike. The NANTEN $^{12}$CO(J=1-0) data reveal molecular clouds positionally associating with the VHE emission, spanning a $\\sim$15~km~s$^{-1}$ range in local standard of rest velocity. } { The VHE/molecular cloud association could indicate a hadronic origin for HESS~J1801$-$233 and HESS~J1800$-$240, and several cloud components in projection may contribute to the VHE emission. The clouds have components covering a broad velocity range encompassing the distance estimates for W~28 ($\\sim$2~kpc) and extending up to $\\sim$4~kpc. Assuming hadronic origin and distances of 2 and 4~kpc for cloud components, the required cosmic-ray density enhancement factors (with respect to the solar value) are in the range $\\sim$10 to $\\sim$30. If situated at 2~kpc distance, such cosmic-ray densities may be supplied by SNRs like W~28. Additionally and/or alternatively, particle acceleration may come from several catalogued SNRs and SNR candidates, the energetic ultra compact HII region W~28A2, and the HII regions M~8 and M~20, along with their associated open clusters. Further sub-mm observations would be recommended to probe in detail the dynamics of the molecular clouds at velocites $>$10~km~s$^{-1}$ and their possible connection to W~28. } ", "introduction": "The study of shell-type supernova remnants (SNRs) at $\\gamma$-ray energies is motivated by the long-held idea that they are the dominant sites of hadronic Galactic cosmic-ray (CR) acceleration to energies approaching the \\emph{knee} ($\\sim 10^{15}$~eV) (e.g. Ginzburg \\& Syrovatskii \\cite{Ginzburg:1}, Blandford \\& Eichler \\cite{Blandford:1}). CRs (hadrons and electrons) are injected into the SNR shock front, and are then accelerated via the diffusive shock acceleration (DSA) process (for a review see Drury \\cite{Drury:2}). Subsequent $\\gamma$-ray production from the interaction of these CRs with ambient matter and/or electromagnetic fields is a tracer of such non-thermal particle acceleration, and establishing the hadronic or electronic nature of the parent CRs in any $\\gamma$-ray source remains a key issue. Two SNRs, RX~J1713.7$-$3946 and RX~J0852.0$-$4622, have so far established shell-like morphology in VHE $\\gamma$-rays (Aharonian \\etal \\cite{HESS_RXJ1713,HESS_VelaJnr,HESS_RXJ1713_II,HESS_VelaJnr_II,HESS_RXJ1713_III}), with spectra extending to 20~TeV and beyond. In particular for RX~J1713.7$-$3946, particle acceleration up to at least 100~TeV is inferred from the H.E.S.S. observations. Although a hadronic origin of the VHE $\\gamma$-ray emission is highly likely in the above cases (Aharonian \\etal \\cite{HESS_RXJ1713_II,HESS_VelaJnr_II}, Berezhko \\& V\\\"olk \\cite{Berezhko:1}, Berezhko, P\\\"uhlhofer \\& V\\\"olk \\cite{Berezhko:2}), an electronic origin is not ruled out. Disentangling the electronic and hadronic components in TeV SNRs may be made easier by studying: (1) SNR $\\gamma$-ray spectra well beyond $\\sim$10 TeV, an energy regime where electrons suffer strong radiative energy losses and due to Klein-Nishina effects the resulting inverse-Compton spectra tend to show a cut-off; (2) older SNRs (age approaching 10$^5$ yr) in which accelerated electrons have lost much of their energy through radiative cooling and do not reach multi-TeV energies; (3) SNRs interacting with adjacent molecular clouds of very high densities $n> 10^3$~cm$^{-3}$. It is the latter regard especially (and to a certain degree the second) which makes the SNR W~28 (G6.4$-$0.1) an attractive target for VHE $\\gamma$-ray studies. In this paper we outline the discovery of VHE $\\gamma$-ray emission from several sites in the W~28 field and briefly discuss their relationship with molecular clouds, W~28, and other potential particle accelerators in the region. W~28 (G6.4$-$0.1) is a mixed-morphology SNR, with dimensions 50$^\\prime$x45$^\\prime$ and an estimated distance between 1.8 and 3.3~kpc (eg. Goudis \\cite{Goudis:1}, Lozinskaya \\cite{Lozinskaya:1}). It is an old-age SNR (age 35000 to 150000~yr; eg. Kaspi \\etal \\cite{Kaspi:1}), thought to have entered its radiative phase of evolution (eg. Lozinskaya \\cite{Lozinskaya:1}) in which much of its CRs have escaped into the surrounding interstellar medium (ISM). We note also that the evolutionary status (Sedov and/or radiative) of shell-type SNRs may depend on the density of their surroundings (see eg. Blondin \\etal \\cite{Blondin:1}). W~28 is distinguished by its interaction with a molecular cloud (Wootten \\cite{Wootten:1}) along its north and northeastern boundaries. This interaction is traced by the high concentration of 1720~MHz OH masers (Frail \\etal \\cite{Frail:2}, Claussen \\etal \\cite{Claussen:1,Claussen:2}), and also the location of very high-density ($n>10^3$~cm$^{-3}$) shocked gas (Arikawa \\etal \\cite{Arikawa:1}, Reach \\etal \\cite{Reach:1}). The shell-like radio emission (Long \\etal \\cite{Long:1}, Dubner \\etal \\cite{Dubner:1}) peaks at the northern and northeastern boundaries where interaction with the molecular cloud is established. Further indication of the influence of W~28 on its surroundings is the expanding HI void at a distance $\\sim$1.9~kpc (Vel\\'azquez \\etal \\cite{Velazquez:1}). The X-ray emission, which overall is well-explained by a thermal model, peaks in the SNR centre but has local enhancements in a region overlapping the northeastern SNR/molecular cloud interaction (Rho \\& Borkowski \\cite{Rho:2}). In the neighbourhood of W~28 are the radio-bright HII regions M~20 (Trifid Nebula at $d \\sim$1.7~kpc Lynds \\etal \\cite{Lynds:1} -- with open cluster NGC~6514), M~8 (Lagoon Nebula at $d\\sim 2$~kpc Tothill \\etal \\cite{Tothill:1} --- containing the open clusters NGC~6523 and NGC~6530) and the ultra-compact HII region W~28A2, all of which are representative of the massive star formation taking place in the region. Further discussion concerning the active star formation in this region may be found in van den Ancker \\etal (\\cite{Ancker:1}) and references therein. Additional SNRs in the vicinity of W~28 have also been identified: G6.67$-$0.42 and G7.06$-$0.12 (Yusef-Zadeh \\etal \\cite{Yusef:1}), G5.55+0.32, G6.10+0.53 and G7.20+0.20 (Brogan \\etal \\cite{Brogan:1}). The pulsar PSR~J1801$-$23 with % spin-down luminosity $\\dot{E} \\sim 6.2\\times 10^{34}$ erg~s$^{-1}$ and distance $d = 13.5$~kpc (based on its dispersion measure) is at the northern radio edge (Kaspi \\cite{Kaspi:1}). More recent discussion (Claussen \\etal \\cite{Claussen:3}) assigns a lower limit of 9.4$\\pm$2.4~kpc for the pulsar distance. W~28 has also been linked to $\\gamma$-ray emission detected at $E>300$~MeV by COS-B (Pollock \\cite{Pollock:1}) and $E>100$~MeV by EGRET (Sturner \\& Dermer \\cite{Sturner:1}, Esposito \\etal \\cite{Esposito:1}, Zhang \\etal \\cite{Zhang:1}). The EGRET source, listed in the 3rd catalogue (Hartman \\etal \\cite{Hartman:1}) as 3EG~J1800$-$2338, is positioned at the southern edge of the radio shell. We have also performed an analysis of EGRET data, with additional data not included in the 3rd catalogue, and results are discussed later in this paper. Previous observations of the W~28 region at VHE energies by the CANGAROO-I telescope revealed no evidence for such emission (Rowell \\etal \\cite{Rowell:1}) and upper limits at the $\\sim$0.2 to 0.5 Crab-flux level for energies $E>1.5$~TeV (1.1 to 2.9$\\times 10^{-11}$~erg~cm$^{-2}$~s$^{-1}$) were set for various regions. ", "conclusions": "\\label{sec:conclusion} In conclusion, our observations with the H.E.S.S. $\\gamma$-ray telescopes have revealed VHE $\\gamma$-ray sources in the field of W~28 which positionally coincide well with molecular clouds. HESS~J1801$-$233 is seen toward the northeast boundary of W~28, while HESS~J1800$-$240 situated just beyond the southern boundary of W~28 comprises three components. Our studies with NANTEN $^{12}$CO(J=1-0) data show molecular clouds spanning a broad range in local standard of rest velocity $V_{\\rm LSR}=$5 to $\\sim$20~km~s$^{-1}$, encompassing the distance estimates for W~28 and various star formation sites in the region. If connected, and at a distance $\\sim$2~kpc, the clouds may be part of a larger parent cloud possibly disrupted by W~28 and/or additional objects related to the active star formation in the region. Cloud components up to $\\sim$4~kpc distance ($V_{\\rm LSR}>$10~km~s$^{-1}$) however, remain a possibility. The VHE/molecular cloud association could indicate a hadronic origin for the VHE sources in the W~28 field. Under assumptions of connected cloud components at a common distance of 2~kpc, or, alternatively, separate cloud components at 2 and 4~kpc, a hadronic origin for the VHE emission implies cosmic-ray densities $\\sim$10 to $\\sim$30 times the local value. W~28 could provide such densities in the case of slow diffusion. Additional and/or alternative particle accelerators such as HII regions representing very young stars, other SNRs/SNR candidates and/or several open clusters in the region may also be contributors. Alternatively, if cloud components at $V_{\\rm LSR}>$10~km~s$^{-1}$ are at distances $d\\sim4$~kpc, as-yet undetected particle accelerators in the Scutum-Crux arm may be responsible. Detailed modeling (beyond the scope of this paper), and further multiwavelength observations of this region are highly recommended to assess further the relationship between the molecular gas and potential particle accelerators in this complex region, as well as the nature of the acclerated particles. In particular, further sub-mm observations (eg. at high CO transitions) will provide more accurate cloud mass estimates, and allow to search for disrupted/shocked gas towards the southern VHE sources. Such studies will be valuable in determining whether or not W~28 and other energetic sources have disrupted molecular material at line velocities $>$10~km~s$^{-1}$." }, "0801/0801.4696_arXiv.txt": { "abstract": "Chaotic inflation predicts a large gravitational wave signal which can be tested by the upcoming Planck satellite. We discuss a SUGRA implementation of chaotic inflation in the presence of moduli fields, and find that inflation does not work with a generic KKLT moduli stabilisation potential. A viable model can be constructed with a fine-tuned moduli sector, but only for a very specific choice of K\\\"ahler potential. Our analysis also shows that inflation models satisfying $\\partial_{i} W_{\\rm inf}=0$ for all inflation sector fields $\\phi_i$ can be combined successfully with a fine-tuned moduli sector. \\vskip 0.1in \\noindent {\\bf Keywords:} inflation, cosmology of theories beyond the SM ", "introduction": "In chaotic inflation models the energy scale of inflation is high, typically of the order of the grand unified scale~\\cite{linde}. As a consequence these models give a large tensor contribution to the density perturbations. This makes them testable by current and future CMB experiments, most notably by the upcoming Planck satellite. However, chaotic inflation is not easy to implement in a supergravity theory~\\cite{kawasaki,kadota}. The inclusion of other high energy physics, such as moduli fields, creates further problems~\\cite{brax,kallosh,kallosh2,kss}. Naturally, any realistic inflation model must be part of some full theory, containing all known physics. The effects of other sectors of the theory on inflation can not be ignored. As shown by Lyth~\\cite{lyth} in the context of slow-roll inflation, a measurable tensor mode requires the inflaton field to change by superplanckian values during inflation. % Examples of such ``large field models'' of inflation are chaotic and natural inflation~\\cite{natural}. At present no string theory derivation of a large field inflaton model exists. The displacement of the inflaton in brane models of inflation is bounded by the size of the compactified space, and in all known models less than the Planck scale~\\cite{baumann,bean}. In all examples of modular inflation the inflationary scale is too low for an appreciable tensor signal~\\cite{kallosh,kallosh2}. N-flation~\\cite{kss,Nflation,grimm}, the stringy realization of assisted inflation~\\cite{assisted}, gives rise to appreciable tensor modes. However, it is not clear whether all of the underlying assumptions are satisfied in these models \\cite{kallosh}. Despite the negative results, so far there is no ``no-go'' theorem stating that string theory cannot give large field inflation. It may very well be that it can be realized in corners of the landscape not yet explored -- after all, the search has only just begun. In this paper we consider a ${\\mathcal N} = 1$ SUGRA implementation of chaotic inflation, and analyse what happens when it is combined with a KKLT-like moduli sector. In our set-up the inflaton and moduli sector only interact gravitationally. Our approach is phenomenological in that we analyse the SUGRA effective field theory, but do not attempt to derive the model from string theory. It should be noted in this context that moduli fields are not unique to string theory. Flat directions abound in any SUSY theory. If SUSY is broken in some hidden sector by non-perturbative physics, the moduli sector has the same qualitative properties as the KKLT model, and our results apply. As mentioned, it is not easy to construct a model of chaotic inflation in the presence of additional moduli fields, even when they are stable. First of all, there is the $\\eta$-problem, present in all models of SUGRA inflation~\\cite{copeland,dine}. The potential during inflation is of the form $V \\sim \\e^K \\tilde V$, which for a canonically normalised inflaton field $\\vp$ gives rise to a large inflaton mass $m_\\vp^2 \\sim H^2$ ruining slow-roll inflation. This can be solved by fine-tuning the K\\\"ahler potential so that the inflaton mass is accidentally small. More elegantly, the inflaton mass can be protected by symmetries. In this paper we will introduce a shift symmetry for the inflaton field that leaves the K\\\"ahler potential invariant to solve the $\\eta$-problem~\\cite{kadota,gaillard,banks}. Inclusion of moduli fields in the system gives rise to a whole new set of obstacles to implement inflation. The moduli fixing potential breaks supersymmetry. Consequently there are soft corrections to the inflaton potential. The soft terms are small in the limit of low scale SUSY breaking, with a small gravitino mass $m_{3/2}^2 \\lesssim H^2$. At the same time, the requirement that the moduli fields remain stabilised in their minimum during inflation, and do not run away to infinity, implies that the moduli masses should be sufficiently large. This requirement is usually expressed as a constraint on the Hubble parameter during inflation $H^2 \\ll m_{\\rm mod}^2$~\\cite{KL}. In a generic potential $m_{\\rm mod}^2 \\sim m_{3/2}^2$ and without fine-tuning (in addition to that required to set the cosmological constant to zero) these requirements are at odds with each other. This is for example the case in the original KKLT model~\\cite{KKLT}. It is difficult to embed large field inflation in such a set up. We see there is tension between keeping the soft corrections to the inflaton potential small, and keeping the moduli fields fixed during inflation. This can be eased if the modulus sector is fine-tuned so that the modulus and gravitino masses are no longer of the same order of magnitude. This is achieved explicitly in the Kallosh-Linde (KL) set-up~\\cite{KL}, which uses a racetrack potential for the modulus field. In this case, parameters are tuned so that the modulus mass is much larger than the gravitino mass. Having the Hubble constant during inflation between these mass scales $m_{\\rm mod}^2 \\ll H^2 \\ll m_{3/2}^2$ offers a way to solve both problems. Note that it also allows the gravitino mass to be in the phenomenologically favoured TeV range, without the need for low scale inflation (in fact, this was the original motivation for KL). In this paper we will analyse chaotic inflation in the presence of a single modulus field with a no-scale K\\\"ahler potential. The models we will study have the superpotential \\be W = W_{\\rm mod}(T) + m \\phi_1 \\phi_2 \\, . \\label{w1} \\ee We consider both a generic KKLT potential and a fine-tuned KL potential. The above inflaton superpotential was first proposed in~\\cite{kawasaki}. Refs.~\\cite{brax,kallosh,kallosh2} added a moduli sector to the set-up. We extend their results by an in-depth discussion of the effects of the moduli dynamics, with an emphasis on finding the conditions for successful inflation. As expected, inflation does not work in the KKLT set-up. Whether KL works depends sensitively on the K\\\"ahler potential for the inflaton fields. Although the moduli corrections are small after inflation due to the fine-tuning in the KL set-up, this is not necessarily true during inflation. During inflation the modulus field $T$ is slightly displaced from its post-inflationary minimum, disrupting the minute fine-tuning of the potential, with potentially large effects. Indeed, consider the following K\\\"ahler potentials \\numparts \\bea K_{1} &=& -3\\log[T+\\bar T] -\\frac12 (\\phi_1 -\\bar{\\phi}_1)^2 + \\phi_2 \\bar{\\phi}_2 \\, , \\label{k1} \\\\ K_{2} &=& -3\\log[T+\\bar T] -\\frac12 (\\phi_1 -\\bar{\\phi}_1)^2 -\\frac12( \\phi_2- \\bar{\\phi}_2)^2 \\, , \\label{k2} \\\\ K_{\\alpha} &=& -3\\log\\left[T+\\bar T- \\frac13 (T+\\bar T)^\\alpha \\, \\phi_2 \\bar{\\phi}_2\\right]-\\frac12 (\\phi_1 -\\bar{\\phi}_1)^2 \\, . \\label{kalpha} \\eea \\endnumparts All K\\\"ahler potentials have a shift symmetry for the inflaton field $\\phi_1$ to solve the $\\eta$-problem. However, as we will show, only $K_{1}$ combined with the KL modulus sector gives a viable model. For all the other models, independent of modular weight $\\alpha$, the coupling between the modulus and inflaton sectors leads to instabilities in the potential, with a runaway behaviour for some of the fields. It is thus crucial to take the dynamics of the modulus field during inflation into account for a correct analysis of the model. This paper is organised as follows. The next section provides the background material, with a concise summary of the KKLT and KL moduli stabilisation potential, as well as a discussion of SUGRA chaotic inflation without moduli. The rest of the paper discusses the combination of chaotic inflation and moduli fields. In section~\\ref{s:model2} we study the model with $K_{2}$~\\eref{k2}. Although inflation does not work, it is useful to analyse why. In section~\\ref{s:model1} we consider the model with $K_{1}$~\\eref{k1}. As mentioned above, this is a viable model of chaotic inflation. We discuss the inflationary predictions, in particular whether the supergravity corrections can leave a signature in the CMB. Finally, in section~\\ref{s:KL} we use the insight gained in the previous sections to discuss more generic combinations of chaotic inflation and KL moduli stabilisation, including models with~\\eref{kalpha}. We end with some concluding remarks. Throughout this article we will work in units with the reduced Planck mass $\\mpl = 1/\\sqrt{8\\pi G_N}$ set to unity. ", "conclusions": "In this paper we studied SUGRA chaotic inflation in the presence of stabilised moduli fields. To avoid the usual $\\eta$-problem a shift symmetry for the inflaton field is introduced. But this is not enough, as the moduli stabilisation sector gives rise to additional contributions to $\\eta$ and $\\epsilon$ which are generically not small. The moduli sector breaks supersymmetry, and as a result the inflaton fields get soft mass contributions of the order of the gravitino mass. These corrections need to be small for successful inflation. But in a generic moduli potential such as KKLT, the modulus mass is of the same order as the gravitino mass, and it is impossible to keep the corrections to the inflaton small while making sure the modulus remains fixed in its minimum during inflation. KL addressed this problem by constructing a fine-tuned moduli potential with $m_{3/2}^2 \\ll m_T^2$. Indeed, calculating the potential in any model in which inflation is combined with a KL moduli sector by adding the respective superpotentials, the moduli corrections to inflation appear small while at the same time the modulus is heavy. All of the above assumes that the modulus $T$ is fixed during inflation. However, the modulus is a dynamical field, and this changes the situation drastically. Although during inflation the modulus is only slightly displaced from its post-inflationary vacuum, this is enough to disrupt the minute fine-tuning of the KL model. The corrections to the effective inflaton potential are generically large, and whether inflation works is a model dependent question. Inflation combined with the KL moduli stabilisation scheme works well if the derivative of the inflaton superpotential during inflation vanishes $(W_{\\rm inf})_i =0$ with $i$ running over all inflaton sector fields. This is for example the case for $D$-term hybrid inflation. On the other hand if $(W_{\\rm inf})_i \\neq 0$, there are large corrections to the masses of the inflaton sector fields, which are missed if the modulus dynamics are not kept. For models with $W_{\\rm inf}$ a polynomial in the shift symmetric inflaton field, these corrections are fatal. If $W_{\\rm inf}$ is some polynomial of inflaton and ``spectator'' fields, the corrections to the $\\eta$-parameter can be harmlessly small if the spectator fields have a small VEV. However, one must also check that the masses of the spectator fields are positive definite during inflation to avoid a run away behaviour. For the chaotic inflation models under consideration this requires the spectator field $\\phi_2$ to have a minimal K\\\"ahler (but note that this model is only ``just'' stable). It is not sufficient for $\\phi_2$ to appear inside the modulus log with unit modular weight, in which case upon a small field expansion it will have a minimal K\\\"ahler. In fact, no matter what the modular weight, if $\\phi_2$ is placed inside the log [see \\eref{Kall}] the spectator field becomes tachyonic during inflation. Our route to a successful inflation model in this paper was to take a specific choice of K\\\"ahler potential that minimises the impact of the moduli corrections. We calculated the inflationary predictions for the viable model 1, which has a minimal kinetic term for the spectator field $\\vp_2$ \\eref{model1}, \\eref{combine}. Although the spectral index $n_s = 0.967$ is the same as for chaotic inflation with a quadratic potential, the values of the slow-roll parameters differ from those of a purely quadratic potential. The difference is largest for those parameters that stabilise $T$ at large values. The degeneracy between the quadratic model and the model with moduli can be broken if tensor perturbations are observed, as this allows us to extract the values of $\\eta$ and $\\epsilon$ from the CMB data. Hence, in the future, with the launch of the Planck satellite, we may be able to observe the presence of moduli fields in the sky. Note that the problems arising from the variation of the modulus $T$ during inflation are not unique to chaotic inflation. Combining moduli with $F$-term hybrid inflation was recently discussed in~\\cite{dp}, where even a careful choice of K\\\"ahler could not save the model. Instead, taking inspiration from~\\cite{anaXW}, the moduli problems were reduced by multiplying the superpotentials of the two sectors, instead of adding them. It would be interesting to see if a similar approach can help chaotic inflation models, although we will leave this for future work. \\ack SCD thanks the Netherlands Organisation for Scientific Research (NWO) for financial support. \\appendix" }, "0801/0801.1258_arXiv.txt": { "abstract": "{ It is widely accepted that the large obliquity of Uranus is the result of a great tangential collision (GC) with an Earth size proto-planet at the end of the accretion process. The impulse imparted by the GC had affected the Uranian satellite system. Very recently, nine irregular satellites (irregulars) have been discovered around Uranus. Their orbital and physical properties, in particular those of the irregular Prospero, set constraints on the GC scenario. }{ We attempt to set constraints on the GC scenario as the cause of Uranus' obliquity as well as on the mechanisms able to give origin to the Uranian irregulars. }{ Different capture mechanisms for irregulars operate at different stages on the giant planets formation process. The mechanisms able to capture the uranian irregulars before and after the GC are analysed. Assuming that they were captured before the GC, we calculate the orbital transfer of the nine irregulars by the impulse imparted by the GC. If their orbital transfer results dynamically implausible, they should have originated after the GC. We then investigate and discuss the dissipative mechanisms able to operate later. }{ Very few transfers exist for five of the irregulars, which makes their existence before the GC hardly expected. In particular Prospero could not exist at the time of the GC. Different capture mechanisms for Prospero after the GC are investigated. Gas drag by Uranus'envelope and pull-down capture are not plausible mechanisms. Capture of Prospero through a collisionless interaction seems to be difficult. The GC itself provides a mechanism of permanent capture. However, the capture of Prospero by the GC is a low probable event. Catastrophic collisions could be a possible mechanism for the birth of Prospero and the other irregulars after the GC. Orbital and physical clusterings should then be expected. }{ Either Prospero had to originate after the GC or the GC did not occur. In the former case, the mechanism for the origin of Prospero after the GC remains an open question. An observing program able to look for dynamical and physical families is mandatory. In the latter case, another theory to account for Uranus' obliquity and the formation of the Uranian regular satellites on the equatorial plane of the planet would be needed. ", "introduction": "Very recently, rich systems of irregular satellites (hereafter irregulars) of the giant planets have been discovered. Enabled by the use of large-format digital images on ground-based telescopes, new observational data have increased the known population of Jovian irregulars to 55 (Sheppard et al. \\cite{shepparda}), the Saturnian population to 38 (Gladman et al. \\cite{gladmanb}, Sheppard et al. \\cite{sheppardb}, \\cite{shepparde}) and the Neptunian population to 7 (Holman et al. \\cite{holman}, Sheppard et al. \\cite{sheppardd}). The Uranian system is of particular interest since a population of 9 irregulars (named Caliban, Sycorax, Prospero, Setebos, Stephano, Trinculo, S/2001U2: XXIV Ferdinand, S/2001U3: XXII Francisco and S/2003U3: XXIII Margaret) has been discovered around Uranus (Gladman et al. \\cite{gladman}, \\cite{gladmana}, Kavelaars et al. \\cite{kavelaars}, Sheppard et al. \\cite{sheppardc}). The discovery of these objects provides a unique window on processes operating in the young Solar System. In the particular case of Uranus, their existence may cast light on the mechanism responsible for its peculiar rotation axis (Parisi \\& Brunini \\cite{parisi}, Brunini et al. \\cite{brunini} (hereafter BP02)). Irregulars of giant planets are characterized by eccentric, highly tilted with respect of the parent planet equatorial plane, and in some case retrograde, orbits. These objects cannot have formed by circumplanetary accretion as the regular satellites but they are likely products of an early capture of primordial objects from heliocentric orbits, probably in association with planet formation itself (Jewitt \\& Sheppard \\cite{jewitta}). It is possible for an object circling about the sun to be temporarily trapped by a planet. In terms of the classical three-body problem this type of capture can occur when the object passes through the interior Lagrangian point, $L_{2}$, with a very low relative velocity. But, without any other mechanism, such a capture is not permanent and the objects will eventually return to a solar orbit after several or several hundred orbital periods. To turn a temporary capture into a permanent one requires a source of orbital energy dissipation and that particles could remain inside the Hill sphere long enough for the capture to be effective. Although currently giant planets have no efficient mechanism of energy dissipation for permanent capture, at their formation epoch several mechanisms may have operated: 1) {\\em gas drag} in the solar nebula or in an extended, primordial planetary atmosphere or in a circumplanetary disk (Pollack et al.\\cite{pollack}, Cuk \\& Burns \\cite{cuk}), 2) {\\em pull-down capture} caused by the mass growth and/or orbital expansion of the planet which expands its Hill sphere (Brunini \\cite{bruninia}, Heppenheimer \\& Porco \\cite{heppenheimer}), 3) {\\em collisionless interactions} between a massive planetary satellite and guest bodies (Tsui \\cite{tsui}) or between the planet and a binary object (Agnor \\& Hamilton \\cite{agnor}), and 4) collisional interaction between two planetesimals passing near the planet or between a planetesimal and a regular satellite. This last mechanism, the so called {\\em break-up} process, leads to the formation of dynamical groupings (e.g. Colombo \\& Franklin 1971, Nesvorny et al. \\cite{nesvornya}). After a break-up the resulting fragments of each progenitor would form a population of irregulars with similar surface composition, i.e. similar colors, and irregular shapes, i.e. light-curves of wide amplitude. Significant fluctuations in the light-curves of Caliban (Maris et al. \\cite{maris}) and Prospero (Maris et al. \\cite{marisa}) and the time dependence observed in the spectrum of Sycorax (Romon et al. \\cite{romon}) suggest the idea of a break-up process for the origin of these bodies. Several theories to account for the large obliquity of Uranus have been proposed. Kubo-Oka \\& Nakazawa (\\cite{kubo-oka}), investigated the tidal evolution of satellite orbits and examined the possibility that the orbital decay of a retrograde satellite leads to the large obliquity of Uranus, but the large mass required for the hypothetical satellite makes this possibility very implausible. An asymmetric infall or torques from nearby mass concentrations during the collapse of the molecular cloud core leading to the formation of the Solar System, could twist the total angular momentum vector of the planetary system. This twist could generate the obliquities of the outer planets (Tremaine \\cite{tremaine}). This model has the disadvantages that the outer planets must form before the infall is complete and that the conditions for the event that would produce the twist are rather strict. The model itself is difficult to be quantitatively tested. Tsiganis et al. (\\cite{tsiganis}) proposed that the current orbital architecture of the outer Solar System could have been produced from an initially compact configuration with Jupiter and Saturn crossing the 2:1 orbital resonance by divergent migration. The crossing led to close encounters among the giant planets, producing large orbital eccentricities and inclinations which were subsequently damped to the current value by gravitational interactions with planetesimals. The obliquity changes due to the change in the orbital inclinations. Since the inclinations are damped by planetesimals interactions on timescales much shorter than the timescales for precession due to the torques from the Sun, especially for Uranus and Neptune, the obliquity returns to small values if it is small before the encounters (Hoi et al. \\cite{hoi}). Large stochastic impacts at the last stage of the planetary formation process have been proposed as the possible cause of the planetary obliquities (e.g. Safronov \\cite{safronov}). The large obliquity of Uranus (98$^\\circ$) is usually attributed to a great tangential collision (GC) between the planet and an Earth-size planetesimal occurred at the end of the epoch of accretion (e.g., Parisi \\& Brunini \\cite{parisi}, Korycansky et al. \\cite{korycansky}). The collision imparts an impulse to Uranus and allows preexisting satellites of the planet to change their orbits. Irregulars on orbits with too large semimajor axis escape from the system (Parisi \\& Brunini 1997), while irregulars with a smaller semimajor axis may be pushed to outer or inner orbits acquiring greater or lower eccentricities depending on the initial orbital elements, the geometry of the impact and the satellite position at the moment of impact. The orbits excited by this perturbation must be consistent with the present orbital configuration of the Uranian irregulars (BP02). In an attempt to clarify the origin of Uranus obliquity and of its irregulars, we are using in this study the most updated information on their orbital and physical properties. In Section 2, we improve the model developed in BP02 for the five Uranian irregulars known at that epoch and extend our study to the new four Uranian irregulars recently discovered by Kavelaars et al. (\\cite{kavelaars}) and Sheppard et al. (\\cite{sheppardc}). The origin of these objects after the GC is discussed in Section 3, where several mechanisms for the origin of Prospero are investigated. The discussion of the results and the conclusions are presented in Section 4. ", "conclusions": "It is usually believed that the large obliquity of Uranus is the result of a great tangential collision (GC) with an Earth-sized proto-planet at the end of the accretion process. We have calculated the transfer of angular momentum and impulse at impact and have shown that the GC had strongly affected the orbits of Uranian satellites. We calculate the transfer of the orbits of the nine known Uranian irregulars by the GC. Very few transfers exist for five of the nine irregulars, making their existence before the GC hardly expected. In particular, Prospero could not exist at the time of the GC. Then, either Prospero had to originate after the GC or the GC did not occur, in which case another theory able to explain Uranus' obliquity and the formation of the Uranian regular satellites would be needed. It is usually believed that the regular satellites of Uranus have accreted from material placed into orbit by the GC (Stevenson et al. \\cite{stevenson}). Within the GC scenario, several possible mechanisms for the capture of Prospero after the GC were investigated. If the Uranian irregulars belong to individual captures and relating the origin of the outer uranian system to a common formation process, gas drag by Uranus' envelope and pull-down capture seem to be implausible. Three-body gravitational encounters might be a source of permanent capture. However, we found that the minimum permanent orbital radius of a guest satellite of Uranus is $\\sim$ 955 $R_U$ while the current semiaxis of Prospero is 645 $R_U$. The GC itself could provide a mechanism of permanent capture and the capture of Prospero could have occurred from a heliocentric orbit as is required within the GC scenario, but due to the low rate of incoming objects it turns out to be difficult. Break-up processes could be the mechanism for the origin of Prospero and the other irregulars in the frame of different scenarios. Prospero might be a fragment of a primary KBO fractured by a collision with another KBO. The fragment could have been captured by Uranus if the two KBOs had a minimum orbital eccentricity of 0.37. Prospero could be a secondary member of a collisional family originated by the collision between another satellite of Uranus and a KBO where the parent satellite of Prospero could have been captured by any mechanism before or after the GC. This process has the disadvantage that it is unlikely that the preexisting satellite were formed from a circumplanetary disk as regular satellites given the large orbital semiaxis required for this object. Since collisional scenarios require in general high collision rates, perhaps the irregulars were originally much more numerous than now. Then, Prospero and also the other irregulars might be the result of mutual collisions among hypothetical preexisting irregulars (Nesvorny et al. \\cite{nesvorny}, \\cite{nesvornyb}) which could have been captured by any other mechanism before the GC. The knowledge of the size and shape distribution of irregulars is important to know their relation to the precursor Kuiper Belt population. It could bring valuable clues to investigate if they are collisional fragments from break-up processes occuring at the Kuiper Belt and thus has nothing to do with how they were individually captured later by the planet, or if they are collisional fragments produced during or after the capture event (Nesvorny et al. \\cite{nesvorny}, \\cite{nesvornyb}). The differential size distribution of the Uranian irregulars approximates a power law with an exponent $q$=1.8 (Sheppard et al.\\cite{sheppardc}). If we assume that the size distribution of the nine irregulars with radii greater than 7 km extends down to radii of about 1 km, we would expect about 75 irregulars of this size or larger (Sheppard et al.\\cite{sheppardc}). The nuclei of Jupiter family comets are widely considered to be kilometer- sized fragments produced collisionally in the Kuiper Belt (Farinella \\& Davis \\cite{farinellaa}). Jewitt et al. (\\cite{jewittb}) compared the shape distribution of cometary nuclei in the Jupiter family with the shape distribution of small main-belt asteroids of similar size (1 km -10 km) and with the shape distribution of fragments produced in laboratory impact experiments. They found that while the asteroids and laboratory impact fragments show similar distribution of axis ratio ($\\mean{b/a}$ $\\sim$ 0.7), cometary nuclei are more elongated ($\\mean{b/a}$ $\\sim$ 0.6). They predict that if comets reflect their collisional origin in the Kuiper Belt followed by sublimation-driven mass loss once inside the orbit of Jupiter, small KBOs should have average shapes consistent with those of collisionally produced fragments (i.e.,$\\mean{b/a}$ $\\sim$ 0.7). To date, constraints on the shapes of only the largest KBOs are available. Prospero being a bit larger than cometary nuclei, displays a variability of 0.21 mag in the R band (Maris et al. \\cite{marisa}). This corresponds to an axis ratio projected into the plane of the sky, b/a of 0.8. The knowledge of the size and shape distribution of irregulars would shed light in the size and shape distribution of small KBOs as well as on the irregulars capture mechanism. Colors are an important diagnostic tool in attempting to unveil the physical status and the origin of the Uranian irregulars. In particular it would be interesting to assess whether it is possible to define subclasses of irregulars just looking at colors, and comparing colors of these bodies with colors of minor bodies in the outer Solar System. Avalilable literature data show a dispersion in the published values larger than quoted errors for each Uranian irregular (Maris et al. \\cite{marisa}, and references therein). We have concluded in Maris et al. (\\cite{marisa}), that the Uranian irregulars are slightly red but they are not as red as the reddest KBOs. An intensive search for fainter irregulars and a long term program of observations able to recover in a self consistent manner light-curves, colors and phase effects informations is mandatory." }, "0801/0801.0859_arXiv.txt": { "abstract": "In this paper, we investigate the structures of galaxies which either have or have had three BHs using $N$-body simulations, and compare them with those of galaxies with binary BHs. We found that the cusp region of a galaxy which have (or had) triple BHs is significantly larger and less dense than that of a galaxy with binary BHs of the same mass. Moreover, the size of the cusp region depends strongly on the evolution history of triple BHs, while in the case of binary BHs, the size of the cusp is determined by the mass of the BHs. In galaxies which have (or had) three BHs, there is a region with significant radial velocity anisotropy, while such a region is not observed in galaxies with binary BH. These differences come from the fact that with triple BHs the energy deposit to the central region of the galaxy can be much larger due to multiple binary-single BH scatterings. Our result suggests that we can discriminate between galaxies which experienced triple BH interactions with those which did not, through the observable signatures such as the cusp size and velocity anisotropy. ", "introduction": "According to recent observations, elliptical galaxies can be classified into two groups: ``weak-cusp'' galaxies and ``strong-cusp'' galaxies \\citep{lauer95, faber97}. The central surface brightness profiles of the weak-cusp galaxies are expressed as $\\Sigma(R) \\propto R^{-\\gamma}$ with $\\gamma \\le 0.3$, and those of the strong-cusp galaxy the same formula with $\\gamma \\ge 0.5$. The slope of the volume density profile of strong-cusp galaxies is around $-2$, being consistent with the isothermal cusp. Such a steep cusp is naturally formed in dissipative process involving gas dynamics and star formation. On the other hand, the weak cusp corresponds to the slope of the volume density shallower than $-1$, which is not likely to be formed through dissipative process. One possible way to form a weak cusp is the merging of two galaxies containing black holes (BHs). When two galaxies, each with a black hole, merge, two BHs sink toward the center of the merger remnant to form a binary through the dynamical friction \\citep{bbr80}. The back reaction of the dynamical friction heats up field stars. As a result, a shallow cusp of stars develops in the central region \\citep{Ebisuzakietal1991,NM99,merritt06}. One problem with this binary BH scenario is what would be the final fate of the binary BH. \\citet{bbr80} pointed out that the merging timescale of the binary BH might be very long, after the binary BH ejected out the stars which can interact with the binary (loss cone depletion). The stars will be supplied in the timescale of the relaxation time, which is much longer than the Hubble time. Recent $N$-body simulations confirmed this theoretical estimate \\citep{MF04, berczik2005}. If galaxies are formed through hierarchical clusterings, in many cases, a binary BH is formed after a merger event. If there were sufficient gas left in the merger, interaction with gas might lead to the quick merging of the two BHs. However, in the case of ``dry'' mergers which would result in the formation of giant ellipticals, by definition not much gas is left and it would be difficult for two BHs to merge. If one galaxy with binary BH and the other with single BH merge, the central BHs form a triple system. Iwasawa, Funato \\& Makino (2006, hereafter referred to as Paper I) investigated the evolution of triple BH system in the galactic center, using $N$-body simulations. They found that the strong binary-single BH interaction \\citep{ME94} and the Kozai cycle \\citep{Kozai62, BLS} drives the eccentricity of the BH binary high enough that two BHs merge quickly through gravitational wave radiation. \\cite{HL07} performed statistical simulations of evolution of central BH systems and reached a similar conclusion. This paper is a follow-up of Paper I. In this paper, we investigate the structure of a galaxy containing three BHs. We also investigated their observational properties which would help us to find galaxies which have (or had) triple BHs. We performed $N$-body simulations of the evolution of triple (or binary) BH in a host galaxy, in order to study the dynamical evolution of the structure of stellar systems containing BHs. In our simulations, both dynamical evolution of BHs and that of field stars are integrated consistently. The interaction between BHs affects not only the evolution of themselves but also the spatial and kinematic structure of field stars around them. In turn, the distribution of field stars affects the interaction between BHs. To understand the structure of galaxies containing BHs, a self-consistent simulation in which the orbits of BHs and stars are treated self-consistently is essential. The structure of this paper is as follows. In section 2, we describe the initial models and the method of our numerical simulations. In section 3, we show the effect of the BH triples dynamics on the structure of the galaxy. Summary and discussion are given in section 4. ", "conclusions": "\\subsection{Possibility to Find Triple BH Systems} In previous sections, we have seen that the density and velocity structure of galaxies with triple BHs is different from that with two BHs. In Paper I, we showed that two of three BHs in a galaxy merge within several dynamical times. The standard hierarchical clustering scenario of galaxy formation suggests that galaxies frequently merge and form ellipticals. Many elliptical galaxies, therefore, might host a binary BH. It is likely that many of them had a triple BH system once upon a time. Furthermore, there may be galaxies with a triple BH system. The lifetime of a triple BH system in a galaxy is several times the dynamical time of the core of the galaxy, which is about $10^8$ years ($100$ M years). If we assume every elliptical galaxy have binary BH, and all large ellipticals are formed through mergings of elliptical galaxy. Speaking, 1\\% of large ellipticals could have triple BH systems now. One possible way to discriminate between a galaxy with binary or single BH and that have (or had) triple BH is the measurement of the cusp radius and density. Our results suggest that the central cusp of a galaxy with three BHs is larger by a factor of few than that with binary BH. Thus the galaxy with a large cusp and low density in the central region might have (or had) three BHs. Another possible way is to use the difference of velocity anisotropy. Anisotropy parameter of galaxy with three BHs is positive (radial) in the outer region due to the heating by BHs. The anisotropy parameter $\\beta$ is estimated by the line of sight velocity and their higher moment, assuming that the galaxy is spherical. However, with this method it is difficult to obtain local anisotropy parameter in the central region of galaxies, unless the central density cusp is steep \\citep{Gerhard93}. Another method is to fit the model with observational data \\citep{Cretton00,Gebhardt00,Gebhardt03}. Figure 10 in \\citet{Gebhardt03} shows anisotropy profile of galaxies. The shapes of some galaxies profile in their paper, for example all ``weak-cusp'' galaxies (NGC3608, NGC4291, NGC4649), are similar to of our result in Figure \\ref{fig9}. These galaxies might contain or have contained triple BHs. \\subsection{Conclusions} In this paper, we investigated the effect of two- or three-BH systems on the structure of galaxies. We found that if the galaxy contains three BHs, (1) multiple three body scattering events reduce the density of the central cusp and increase the cusp radius and (2) in outer region, orbits of stars are likely to be radial. These difference allow us to discriminate the galaxies with two or three BHs." }, "0801/0801.2547_arXiv.txt": { "abstract": "% {} {We investigate the present-day photometric properties of the dwarf spheroidal galaxies in the Local Group. From the analysis of their integrated colours, we consider a possible link between dwarf spheroidals and giant ellipticals. From the analysis of the $M_{V}$ vs (B-V) plot, we search for a possible evolutionary link between dwarf spheroidal galaxies (dSphs) and dwarf irregular galaxies (dIrrs). } {By means of chemical evolution models combined with a spectro-photometric model, we study the evolution of six Local Group dwarf spheroidal galaxies (Carina, Draco, Sagittarius, Sculptor, Sextans and Ursa Minor). The chemical evolution models, which adopt up-to-date nucleosynthesis from low and intermediate mass stars as well as nucleosynthesis and energetic feedback from supernovae type Ia and II, reproduce several observational constraints of these galaxies, such as abundance ratios versus metallicity and the metallicity distributions. The proposed scenario for the evolution of these galaxies is characterised by low star formation rates and high galactic wind efficiencies.}{ Such a scenario allows us to predict integrated colours and magnitudes which agree with observations. Our results strongly suggest that the first few Gyrs of evolution, when the star formation is most active, are crucial to define the luminosities, colours, and other photometric properties as observed today. After the star formation epoch, the galactic wind sweeps away a large fraction of the gas of each galaxy, which then evolves passively. \\\\ Our results indicate that it is likely that at a certain stage of their evolution, dSphs and dIrrs presented similar photometric properties. However, after that phase, they evolved along different paths, leading them to their currently disparate properties. }{} ", "introduction": "The dwarf spheroidal galaxies (dSphs) are among the most common types of galaxies in the universe. They are found normally in groups (C\\^ot\\'e et al. 1997) and clusters (Phillips et al. 1998, Ferguson \\& Sandage 1991). The ones found in the Local Group have become an increasingly important matter of study in the last few years due to their proximity, which enables one to study in detail objects and processes which were formally restricted to our own Galaxy. Several issues regarding the formation and evolution of the Local Group dSphs can help in the attempt to clarify the whole subject of galaxy formation. For example, how did the dSphs form? When? What mechanism ruled their evolution? Are these galaxies remnants of the building blocks from which larger galaxies assembled? Is their evolution mainly affected by the environment or do internal processes play a major role? Are these galaxies linked to other types of dwarf galaxies in the context of any evolutionary scenario? In order to answer these questions, one should try to understand not only the present day properties of the dSphs, but also try to understand their past evolutionary history. One possible procedure is to make use of models which, being based on present day observational constraints, allow one to trace the past evolution of the dSphs. Originally, the local dSphs were believed to be very old simple systems, similar to globular clusters (Shapley 1938). However, more recent deep photometric observations of main sequence turn-off stars revealed also intermediate-age populations, as well as a significant metallicity range, different from most globular clusters. In fact, analysis of colour-magnitude diagrams (CMDs) suggested that these galaxies are characterised by complex and different star formation (SF) histories (van den Bergh 1994; Hernandez, Gilmore, Valls-Gabaud 2000; Dolphin et al. 2005). In almost all cases, the mechanisms which trigger and control the SF are yet unknown and several scenarios have been proposed. These scenarios should also explain the complete lack of gas in these galaxies, the low metallicities, the low values of [$\\alpha$/Fe] relative to Galactic stars with the same [Fe/H] (Bonifacio et al. 2000; Shetrone, C\\^ot\\'e, Sargent 2001; Shetrone et al. 2003; Tolstoy et al. 2003, Bonifacio et al. 2004; Venn et al. 2004; Sadakane et al. 2004; Geisler et al. 2005; Monaco et al. 2005) and the metallicity distributions, including the large metallicity range (Koch et al. 2005, Bellazzini et al. 2002). \\\\ Several attempts to model the properties of dSphs have been performed, generally following different approaches. By means of high-resolution cosmological numerical simulations, Ricotti \\& Gnedin (2005) and Kawata et al. (2006) studied dSphs in a cosmological framework. Their studies are useful to understand the link between the dSphs and the first galaxies, but cannot investigate in detail the chemical properties of dSphs and the relative roles of SNe Ia and II in the chemical enrichment and gas ejection processes. Marcolini et al. (2006), by means of a chemo-dynamical evolution model, studied the evolution of the interstellar medium (ISM) of the Draco dSph. In their picture no galactic wind develops and, in order to deplete the galaxy of its gas and to stop star formation, one must invoke an external mechanism, such as ram pressure stripping or tidal interactions. Fenner et al. (2006) and Ikuta $\\&$ Arimoto (2002) studied the properties of dSphs by means of chemical evolution models including galactic winds. Their results indicate that a small fraction of the ISM is carried away by the SN-driven winds and confirm the need of external gas removing processes to reproduce the present-day gas fractions of dSphs (see also Mori \\& Burkert 2000, Mayer et al. 2006). Lanfranchi $\\&$ Matteucci (2003, 2004, hereinafter LM03, LM04), alternatively, by means of a detailed chemical evolution model with galactic winds, were able to reproduce many observational constraints of six local dSph galaxies (Carina, Draco, Sagittarius, Sculptor, Sextans, Ursa Minor). The scenario proposed by these authors considered low efficiency star formation rates (SFR) derived from colour-magnitude diagrams together with intense galactic winds. \\\\ LM03,04 were able to reproduce not only the chemical properties, but also the lack of central neutral gas and the metallicity distributions of the studied galaxies. In particular, the [$\\alpha$/Fe] vs [Fe/H] and neutron capture element ratios in these galaxies are well reproduced as well as the stellar metallicity distribution (SMD).\\\\ The photometric properties of the local dSphs, however, are rarely addressed in any of these models, even though they could help, not only in constraining the formation and evolution of these galaxies, but also in clarifying the subject of a possible evolutionary connection between the gas poor dSph galaxies and gas rich dwarf irregular galaxies (dIrrs). In such a scenario, a starburst in a dIrr gives rise to a super wind which removes all the gas of the galaxy (which could be removed also by ram pressure stripping or tidal stripping) and halts the SF, giving rise to a dSph galaxy (Lin $\\&$ Faber 1983; Dekel $\\&$ Silk 1986; van den Bergh 1994; Papaderos et al. 1996; Davies $\\&$ Phillipps 1988). There are, however, several drawbacks in that scenario, from both the chemical and photometric point of view. First, the large scale distribution of dSphs is substantially different from that of dIrrs. Most of the dSphs cluster around the two giant spirals of the Local Group (Grebel 1998), with very few exceptions. On the other hand, most of the dIrrs lie at large ($>500 kpc$) distances from the large galaxies (Mateo 1998). This is probably linked to the morphology-density relation for dwarf galaxies and is interpreted as an evidence of environmental effects on galaxy evolution (Grebel 2001). In particular, it may be possible that their proximity to large galaxies had some effects on the evolution of dSphs. \\\\ Futhermore, the dIrrs and dSphs show rather different observational properties. The integrated colour-magnitude diagram of the Local Group dwarf galaxies exhibits a clear distinction between the dSphs and the dIrrs, which occupy different regions (Mateo 1998). A few galaxies, namely transition type dwarfs, exhibit intermediate photometric properies between those of dSphs and dIrrs. Also the luminosity/metallicity relation can be used to distinguish between dIrrs and dSphs at the present epoch (Mateo 1998, Grebel et al. 2003). At the same luminosity, the dIrrs tend to exhibit lower metallicities than the dSph galaxies. These two facts could be related to different evolutionary histories for the two types of dwarfs and only a study of their past history could help in clarifying this issue. In this work we show that a spectro-photometric code coupled with a chemical evolution model allows us to investigate the past evolution of the dSphs and impose constraints on the formation and evolution of these galaxies and also to the possible connection with other types of dwarf and giant galaxies. A chemical evolution code which reproduces successfully several observational constraints can provide the parameters required to predict the evolution of the photometric properties of a sample of six local dSph galaxies. The paper is organized as follows: in Sect. 2, the chemical evolution models which reproduce the chemical data of these galaxies and their results are described. In Sect. 3 we describe the spectro-photometric code. The results of our models compared to observational data are shown in Sect. 4 and finally in Sect. 5 we draw some conclusions. ", "conclusions": "By means of a chemical evolution model combined with a spectro-photometric code we were able to predict the evolution of several photometric properties of six Local Group Dwarf Spheroidal galaxies (Carina, Draco, Sagittarius, Sculptor, Sextans and Ursa Minor). The chemical evolution models adopt up-to-date nucleosynthesis from intermediate mass stars, massive stars and SNe type II and Ia and take into account the contribution of SNe to the energetics of the ISM. For the six dSphs, the star formation histories are taken from the observed colour-magnitude diagrams (CMDs, Dolphin et al. 2005). The proposed scenario for the evolution of these galaxies is characterised by low star formation rates and high galactic wind efficiencies. Such a scenario allows us to predict colours and magnitudes in agreement with observations. The main conclusions can be summarized as follows: \\begin{itemize} \\item the total luminosities of 5 (Sculptor, Carina, Sextans, Ursa Minor and Draco) out of 6 dSphs galaxies analysed here are dominated by the SSPs formed at the lowest metallicities, i.e. $Z\\le 0.0004$. This is a consequence of the low SF efficiency and of the intense wind which removes a large fraction of the gas of the galaxy, almost halting the SF and preventing metal-rich stars to be formed. For each galaxy, the stellar populations dominating the metallicity distribution (LM04) dominate also the total light. \\item in the case of Sagittarius, the B and K luminosities are dominated by stellar populations with higher metallicities, i.e. $Z\\ge 0.004$, due to the much higher SF efficiency adopted for this galaxy ($>$ 10 times higher than the ones of the other dSphs); \\item only during the SF phase, when all galaxies exhibit a high gas content, the effects of dust extinction in the total luminosities in the U, V, B bands are noticeable. After the onset of the galactic winds and the consequent removal of the interstellar gas, the effects of extinction become negligible; \\item We compared the predicted current integrated (U-V) and (V-K) colours as a function of the absolute V magnitude for the dSphs to the colour-magnitude relation for giant spheroids in clusters. The main differences between dSphs and the giant ellipticals concern their star formation histories. In particular, the formation of dSphs occurred with a lower degree of synchronicity than the formation of giant ellipticals. The differences may also be due to environmental effects, since clusters are environments denser than the Local Group. In the colour-magnitude plots, the six local dSphs studied here seem to form a sequence similar to the one observed for their counterparts in clusters, although with a slightly steeper slope and a larger dispersion in their ages. Although we can not draw any firm conclusion on this aspect, it is not unlikely that the dSphs can be regarded as the lowest-mass tails of their larger counterparts and not the building blocks from which they were assembled. This issue represents a challenge to all galaxy formation models based on the popular $\\Lambda$ cold dark matter cosmological scenario. \\item The study of the evolution of the integrated V magnitude as a function of the (B-V) colour suggests that, during their past history, the dSphs may have shown photometric properties similar to the ones of the present-day dIrrs. It is possible that dSphs and dIrrs shared a common progenitor phase, and then evolved through different paths. After the common phase, dIrrs have experienced little evolution, mantaining most of their gas and blue colours. dSphs have instead experienced a stronger evolution, losing all of their gas and showing present-time colours redder than dIrrs. Transition types might represent a sub-class of dIrrs, which may evolve into dSphs. \\end{itemize}" }, "0801/0801.2771_arXiv.txt": { "abstract": "Void regions of the Universe offer a special environment for studying cosmology and galaxy formation, which may expose weaknesses in our understanding of these phenomena. Although galaxies in voids are observed to be predominately gas rich, star forming and blue, a sub-population of bright red void galaxies can also be found, whose star formation was shut down long ago. Are the same processes that quench star formation in denser regions of the Universe also at work in voids? We compare the luminosity function of void galaxies in the 2dF Galaxy Redshift Survey, to those from a galaxy formation model built on the Millennium Simulation. We show that a global star formation suppression mechanism in the form of low luminosity ``radio mode'' AGN heating is sufficient to reproduce the observed population of void early-types. Radio mode heating is environment independent other than its dependence on dark matter halo mass, where, above a critical mass threshold of approximately $M_{\\rm vir}\\!\\sim\\!10^{12.5} M_\\odot$, gas cooling onto the galaxy is suppressed and star formation subsequently fades. In the Millennium Simulation, the void halo mass function is shifted with respect to denser environments, but still maintains a high mass tail above this critical threshold. In such void halos, radio mode heating remains efficient and red galaxies are found; collectively these galaxies match the observed space density without any modification to the model. Consequently, galaxies living in vastly different large-scale environments but hosted by halos of similar mass are predicted to have similar properties, consistent with observations. ", "introduction": "how does one understand early-type galaxies in voids when void environments are typically very gas rich and slowly evolving, which tends to promote star formation rather than suppress it. At this point it may be valuable to backtrack somewhat and revisit the two main mechanisms acting in the model that can turn blue galaxies red. The first is important for satellite (i.e. non-central) galaxies. Upon infall into a more massive system, any extended hot gas around the (now) satellite is stripped by the denser medium and added to the more massive halo. Such ``strangulation'' drives a satellite galaxy to redden rapidly once its remaining cold disk gas is exhausted. The second is the radio mode heating discussed above and in Section~\\ref{sec:model}. Radio mode AGN operate only in central galaxies where the host halo has grown above a critical mass threshold, approximately $M_{\\rm vir} \\sim 10^{12.5} M_\\odot$. So what physical processes have occurred to produce in the observed 2dFGRS early-type void population? It may be that such red galaxies are simply a population of satellites and have survived long enough for an effect on their colours to be seen. To test this idea using the model we perform a simple and revealing exercise. From our knowledge of which galaxies are central and which are satellites, we remove those that are satellites and hence those that could have experienced the strangulation effect. This is shown by the dotted lines in Figure~\\ref{fig:LFcol}. Although satellites do contribute to the void early-type luminosity function at $M_{\\rm b_J}\\!-\\!5\\log_{10}\\! h \\simgt -18.5$, they are clearly a minor component brighter than this. Indeed, all bright early-type void galaxies are centrals in our galaxy formation model. The second of our star formation shut-down mechanisms, i.e. radio mode low luminosity AGN heating, requires that central galaxies reside in sufficiently large dark mater halos. Such halos are not expected to be common in void regions of the Universe where low mass halos dominate. In low mass halos the central density of hot gas is not high enough to maintain the low Eddington accretion needed to power a central AGN outflow. We check this explicitly in Figure~\\ref{fig:MF}, where we plot the halo mass function for halos in all environments (left panel) and void environments (right panel) (wide-dashed lines in both). In addition, we break the halo mass functions into those halos hosting red central galaxies (solid lines) and those hosting blue central galaxies (dashed lines). Figure~\\ref{fig:MF} reveals the origin of the void early-type population -- they are simply central galaxies that live in halos massive enough to have the low luminosity radio mode operating. In other words, the physics assumed by the model that transforms blue galaxies into red operates independent of environment but can still produce the observed environmental trends in voids (note that the transition from blue-dominated to red-dominated dark matter halos begins at approximately the same mass in both panels, $M_{\\rm vir} \\sim 10^{12} M_\\odot$). Specifically, \\emph{it is the shift in halo mass function with environment that changes, not the transformation mechanism itself.} This picture has important consequences for galaxy evolution more generally. First, it indicates the necessity for active galaxies in voids, confirmed in a number of studies (see, e.g., \\citealt{Constantin2008} at $z\\sim0$, and \\citealt{MonteroDorta2008} at $z\\sim1$). Second, it requires that void galaxies should have similar properties to those in all other environments, assuming the galaxies compared are hosted by halos of similar mass. This is a theoretical confirmation of numerous observational studies that have shown that environments on scales larger than a few Mpc (i.e. outside the dark matter halo) are not important for galaxy formation \\citep[e.g. see][and references therein]{Blanton2007}. For example, \\cite{Patiri2006} investigate the properties of void galaxies in the Sloan Digital Sky Survey and compare to those of a similarly constructed mock catalogue using the same \\cite{Croton2006} model used here. They found little difference in the mean properties of void galaxies relative to the field, specifically in colour, specific star formation rate, and morphology. Theoretically one may expect large-scale environment to be important, e.g. the effect of assembly bias for dark matter halos \\citep{Gao2005}. However the galaxy population does not appear to be overly sensitive to this \\citep[see, for example,][]{Croton2007}. A similar conclusion was recently arrived at in a complimentary analysis by \\cite{Tinker2008}. It has been suggested that void environments pose a problem for galaxy formation theory in a $\\Lambda$CDM universe due to the apparent over-abundance of dark matter halos in voids relative to the observed abundance of galaxies \\citep{Peebles2001}. In this paper we show that no conflict exists within the current observational uncertainty. \\begin{itemize} \\item Our model can reproduce the observed abundance of void galaxies, both globally and by colour, without any modification or additional environment dependent physics. \\item Early-type ``red and dead'' red sequence galaxies appear naturally in the voids. They arises because of a shift in the halo mass function in low density environments combined with an environment independent star formation shut-down mechanism efficient above a critical halo mass (here, radio mode AGN). Together, these approximately produce the correct observed abundance. \\item Some notable consequences follow from our results. For example, at a given host halo mass, void galaxies are expected to have similar properties on average to those in the field, because such galaxies will have had similar evolutionary histories to field and cluster galaxies. \\end{itemize} Voids and void galaxies provide an important probe of both cosmology and galaxy formation. Future wide-field surveys will produce large-scale maps of the Universe out to high redshift, from which void evolution can be studied in detail. Such work will further constrain galaxy formation theory, both in under-dense and the wider field environment.", "conclusions": "\\label{sec:results} \\subsection{The void luminosity function} \\label{subsec:LF} \\begin{figure*} \\plotscaled{./figures/figure3_new.ps} \\caption{The Millennium Simulation halo mass function (left panel) and halo mass function in voids (right panel). In both panels the halo mass function is broken up into those who host red central galaxies (solid lines) and those that host blue central galaxies (dashed lines). Red sequence galaxies occupy the most massive halos in all environments -- these halos are subject to the radio mode low luminosity AGN heating that ultimately shuts down star formation. Notably, this shutdown begins at approximately the same mass in both panels, $M_{\\rm vir} \\sim 10^{12-12.5} M_\\odot$. This implies that the (environment independent) radio mode heating, plus a shift in the halo mass function with environment, is sufficient to reproduce the observed abundance of void early-type galaxies seen in Figure~\\ref{fig:LFcol}.} \\label{fig:MF} \\end{figure*} We begin with Figure~\\ref{fig:LFall}, which shows the luminosity function of model galaxies for both the population as a whole, and all galaxies in void regions. Here we define a void as \\cite{Croton2005} did, $\\delta_8\\!<\\!-0.75$, i.e. where the density within an \\eightMP\\ radius of the galaxy is less than $25\\%$ the mean. Over-plotted are the observational results for the full 2dFGRS \\citep[filled circles,][]{Norberg2002} and void galaxy luminosity function \\citep[open squares,][]{Croton2005}. As was the focus of \\cite{Croton2006}, the global luminosity distribution of model galaxies is a good match to the local observations. Important model ingredients that produced this result are the supernova in shallower potentials that expel disk gas to reduce star formation and flatten the faint-end slope, and the inclusion of an AGN heating source in large halos to starve massive central galaxies of star forming fuel, resulting in the observed bright-end exponential cut-off, as described in Section~\\ref{sec:model}. Without these two critical aspects in the model, the luminosity function instead would have a power-law shape reflecting the underlying mass function of halos \\citep{Benson2003}. Note that there remain discrepancies in the top curve of Figure~\\ref{fig:LFall}, notably that the model over-predicts the abundance of very bright galaxies. As discussed in \\cite{Croton2006}, these galaxies represent a population undergoing strong star formation and merger-induced starbursts. In such ultraluminous infrared galaxies (ULIRGs), nearly all the light from young stars is absorbed by dust and re-radiated in the mid- to far-infrared \\citep{Sanders1996}. Improved modelling of the effects of dust are required to adequately reproduce the properties of such systems. The lower solid line in Figure~\\ref{fig:LFall} shows the model result for the void galaxy population in the Millennium Simulation. The good overall agreement is a result of the physical prescriptions and global parameter choices in the \\cite{Croton2006} model and not additional fine-tuning (of which there was none). There is some over-prediction of the very brightest void galaxies. However the discrepancy is not significant enough to expect that improvements to existing aspects of the model, e.g. dust as described above, cannot alleviate such differences. Additionally, we find a small over-prediction of faint ($M_{\\rm b_J}\\!-\\!5\\log_{10}\\! h \\simgt -18.5$) galaxies. Unfortunately, systematic variations in the observed faint-end galaxy luminosity function exist at a level greater than this (e.g. contrast \\citealt{Cole2001} to \\citealt{Huang2003}). Figure~\\ref{fig:LFall} alone answers the challenge posed by Peebles and outlined in the Introduction. Specifically, in a $\\Lambda$CDM Universe the distribution of dark matter halos in voids, coupled with a realistic model of galaxy formation, is consistent with voids as observed in the real Universe. Such regions are typically not devoid of galaxies as claimed in \\cite{Peebles2001}. The remainder of this letter will aim to dissect and understand the good agreement shown in Figure~\\ref{fig:LFall} between model and observation. \\subsection{Galaxy colours in voids} \\label{subsec:col} The colour of a galaxy tells us a lot about the relevant physics that has been dominant during its evolution. A red spectrum typically indicates that a star formation shutdown mechanism has been operating for a significant part of the galaxy's lifetime. We can use this knowledge, in conjunction with our theoretical model, to gain insight into how shutdown may occur and to what degree it may (or may not) be environment specific. To this end, in Figure~\\ref{fig:LFcol} we again plot the void galaxy luminosity function for both the 2dFGRS \\cite[symbols,][]{Croton2005} and semi-analytic model (lines), but now broken up by galaxy spectral type (early/late) or colour (red/blue). Early-type (triangles) and late-type (squares) 2dFGRS galaxies are determined using the principal component analysis of \\cite{Madgwick2003}. For the model we use the bi-modal galaxy colours to separate red (solid line) from blue (dashed line) at $m_{b_{\\rm J}}\\!-\\!m_{r_{\\rm F}}\\!=\\!1.07$ \\citep[see figure~9 of][]{Croton2006}. Comparing model to observation, Figure~\\ref{fig:LFcol} shows agreement (to within $2\\sigma$) between the two sets of early-type/red and late-type/blue void luminosity functions. This has occurred as a byproduct of matching the model to the observed \\emph{global} properties and only these -- no special void environment physics was needed. Adding detail to the physical prescriptions assumed by the model would presumably improve the agreement further, but this is not our focus here. The early-type 2dFGRS void galaxies in Figure~\\ref{fig:LFcol} return" }, "0801/0801.4294_arXiv.txt": { "abstract": "We describe the results of a radio continuum survey of the central 4\\sdeg$\\times$1\\sdeg\\ with the 100 m Green Bank Telescope (GBT) at wavelengths of 3.5, 6, 20, and 90 cm. The 3.5 and 6 cm surveys are the most sensitive and highest resolution single dish surveys made of the central degrees of our Galaxy. We present catalogs of compact and extended sources in the central four degrees of our Galaxy, including detailed spectral index studies of all sources. The analysis covers star-forming regions such as Sgr B and Sgr C where we find evidence of a mixture of thermal and nonthermal emission. The analysis quantifies the relative contribution of thermal and nonthermal processes to the radio continuum flux density toward the GC region. In the central 4\\sdeg$\\times$1\\sdeg\\ of the GC, the thermal and nonthermal flux fractions for all compact and diffuse sources are 28\\%/72\\% at 3.5 cm and 19\\%/81\\% at 6 cm. The total flux densities from these sources are $783\\pm52$ Jy and $1063\\pm93$ Jy at 3.5 and 6 cm, respectively, excluding the contribution of Galactic synchrotron emission. ", "introduction": "The central few hundred parsecs of the Milky Way comprise a region in the Galaxy unique for its high stellar density, intense ionizing radiation field, massive black hole, enhanced density of cosmic rays, and unusual magnetized structures \\citep{f04,r01,y07,b87,y04}. The extent of the GC region is roughly 400 pc in diameter, defined by a region with relatively high gas density \\citep[$n_{\\rm{H_2}}\\gtrsim10^4$ cm$^{-3}$;][]{h98}. This region, sometimes called the Galactic nucleus or ``central molecular zone'', produces 5\\%--10\\% of the Galaxy's infrared and Lyman continuum luminosity and contains 10\\% of its molecular gas \\citep{b87,m96}. The density and physical diversity of objects in the GC region make it highly complex and require a wide range of observations to unravel. At the simplest level, it is important to understand how the basic components of the region interact. Nonthermal emission, in the form of supernova remnants (SNRs) and nonthermal radio filaments \\citep[NRFs;][]{y84}, dominates the cm-wavelength emission in the region \\citep[e.g.,][]{a79,d95}. The relativistic component of the ISM has also been observed at TeV energies as a diffuse source tracing the molecular gas distribution \\citep{a06}. However, it is not clear how these energetic electrons traced by the nonthermal emission affect other components of the GC interstellar medium. For example, energetic electrons may explain the unusually high ambient gas temperature in the region and subsequent low star formation efficiency \\citep{y07}. Thermal emission traces star formation regions and photoionized clouds. An accurate census of thermal emission can constrain the amount of star formation occurring there or find new star forming regions. Separating the thermal and nonthermal processes also allows an estimate of their basic properties \\citep[$n_e$, $B$;][]{h96}. Single-dish radio continuum observations are a useful tool for studying the large-scale properties of the GC region. Several other single-dish surveys of radio continuum emission from the GC region have been conducted \\citep{a79,ha87,r90,h92,d95}. \\citet{a79} observed the Galactic disk from $l=60$\\sdeg\\ to the GC region near 6 cm with the 100 m Effelsberg telescope. This was one of the first single-dish surveys of the GC region with a resolution of a few (2\\damin6) arcminutes and it discovered many compact and diffuse sources. Since that time, the Parkes 64 m and Effelsberg 100 m telescopes have been used to create complete surveys of the Galactic disk in the northern and southern celestial skies near 12 cm, which were sensitive to faint emission on large scales \\citep{r90,d95}. These studies revealed dozens of new SNR candidates, compact \\hii\\ regions, and Galactic loops and spurs, vividly demonstrating the chaotic structure of the Galactic interstellar medium. Although most of the world's best radio telescopes have surveyed the radio continuum in the GC region, there has been limited study of spectral indices on arcminute scales. We were motivated to extend upon previous observations using the largest, fully-steerable telescope, the Green Bank Telescope\\footnote{The GBT is operated by the National Radio Astronomy Observatory, which is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.} (GBT). The high resolution and sensitivity of the GBT observations allow us to separate and quantify flux from all thermal and nonthermal emitters. Section \\ref{gcsurvey_obs} describes the observations and data reduction. In \\S\\ \\ref{gcsurvey_res}, the results of the survey are described, including the compilation of compact and extended source catalogs at 3.5 and 6 cm and a detailed discussion of the radio continuum properties for each source in the region. We calculate the percentage of flux from sources in the central degrees of the Galaxy from thermal/nonthermal processes at 3.5 and 6 cm. Finally, \\S\\ \\ref{gcsurvey_con} summarizes the results of this analysis. A study of the continuum emission from the GC lobe \\citep{s84} is not included here, but will be discussed in detail in a later paper \\citep{l08}. ", "conclusions": "\\label{gcsurvey_con} This paper has shown results from a new survey of the radio continuum emission from the central degrees of the Galaxy at 90, 20, 6, and 3.5 cm with the GBT. The 6 and 3.5 cm surveys are the most sensitive, highest resolution, single-dish radio surveys of the central degrees of the GC made at these wavelengths. The primary products of this study are catalogs of all compact and extended sources, including a spectral index analysis of these sources. We have shown that the compact sources ($\\theta\\lesssim5$\\arcmin) detected at 6 and 3.5 cm surveys are most likely to be Galactic \\hii\\ regions and are mostly found near well-known \\hii\\ complexes, such as Sgr B and Sgr E. About one quarter of the sources detected at 3.5 cm are also detected at 6 cm; most of these sources have thermal spectral indices. Extended, nonthermal emission in the 3.5, 6, and 20 cm surveys is found associated with GC star-forming regions. The emission is on size scales of tens of arcminutes and is particularly found for $l=358\\ddeg5-359\\ddeg5$. The brightness-distribution function of NRFs estimated by high resolution observations indicates that they are much more numerous at low flux densities, so the extended nonthermal emission may yet be resolved as new, nonthermal radio filaments. Nonthermal emission also traces supernova remnants, and these observations find nonthermal emission in Sgr B, consistent with a recent report of a supernova in that star-forming region. Another interesting result from our spectral index study is that the 6/3.5 cm spectral index distribution around the G359.1--0.5 SNR is consistent with the idea that it is interacting with a neighboring molecular cloud that may be in the GC region. We also find a structure south of the Arched filaments that is thermal and seems to be morphologically connected to the well-known Arched filaments. The Arched filaments and the new southern Arched filaments (G0.07--0.2) form a contiguous structure that is a radio-continuum counterpart to the Radio Arc Bubble. The Bubble surrounds the brightest part of the nonthermal emission from the Radio Arc and is close to the dense star clusters, the Quintuplet and Arches clusters. Several recent works has definitively shown that the ionization of the Arched filaments is caused by the Arches star cluster, while the rest of the Bubble is ionized by the Quintuplet star cluster \\citep{r01,la02,si07}. \\citet{si07} further suggests that the Bubble was formed by the winds and ionizing photons of massive stars the Quintuplet cluster. This work highlights the fact that the Bubble --- a thermal structure formed by stellar action --- has a clear effect on the brightness of the nonthermal Radio Arc. The reason for the enhanced brightness inside the Bubble is not clear, but should be considered in models for the formation of NRFs. Through a combination of slice and integrated flux analysis, all objects detected in the surveys are catagorized as having a thermal or nonthermal origin. Thus, the distribution of flux between thermal and nonthermal sources can be quantified. Within the central 4\\sdeg$\\times$1\\sdeg\\ of the Galaxy, the thermal to nonthermal flux fractions for all discrete emission are 28\\%/72\\% at 3.5 cm and 19\\%/81\\% at 6 cm. This does not include the background synchrotron contribution from the Galactic plane, which begins to dominate the Galaxy's flux density for wavelengths longer than 6 cm. Also, some of these sources in the field are likely to be in the foreground of the GC region, although the density of gas and stars is generally much higher in the GC region, so most sources are likely to truly be in the central few hundred parsecs of the Galaxy. The high fraction of nonthermal emission in the radio continuum emission is consistent with the idea that the cosmic ray density is enhanced in the nuclear disk, assuming that the magnetic field is not unusually strong ($\\sim1$ mG) in the region. Other studies have also found evidence for such a cosmic ray enhancement, which seems heat and ionize molecular gas in the GC region, possibly affecting the star formation process \\citep{y07}." }, "0801/0801.3682_arXiv.txt": { "abstract": "{}% {% We present the current status of ongoing searches for molecular hydrogen in high-redshift ($1.8<\\zabs \\le 4.2$) Damped Lyman-$\\alpha$ systems (DLAs) capitalising on observations performed with the ESO Very Large Telescope (VLT) Ultraviolet and Visual Echelle Spectrograph (UVES). }% {% We identify 77 DLAs/strong sub-DLAs, with $\\log N(\\HI)\\ge 20$ and $\\zabs >1.8$, which have data that include redshifted H$_2$ Lyman and/or Werner-band absorption lines. This sample of H\\,{\\sc i}, H$_2$ and metal line measurements, performed in an homogeneous manner, is more than twice as large as our previous sample (Ledoux et al. 2003) considering every system in which searches for H$_2$ could be completed so far, including all non-detections. }% {% H$_2$ is detected in thirteen of the systems, which have molecular fractions of values between $f\\simeq 5\\times 10^{-7}$ and $f\\simeq 0.1$, where $f=2N($H$_2)/(2N($H$_2)+N(\\HI))$. Upper limits are measured for the remaining 64 systems with detection limits of typically $\\log N($H$_2)\\sim 14.3$, corresponding to $\\log f<-5$. We find that about 35\\% of the DLAs with metallicities relative to solar [X/H$]\\ge -1.3$ (i.e., 1/20$^{\\rm th}$ solar), with X~=~Zn, S or Si, have molecular fractions $\\log f>-4.5$, while H$_2$ is detected -- regardless of the molecular fraction -- in $\\sim 50$\\% of them. In contrast, only about 4\\% of the [X/H$]<-1.3$ DLAs have $\\log f>-4.5$. We show that the presence of H$_2$ does not strongly depend on the total neutral hydrogen column density, although the probability of finding $\\log f>-4.5$ is higher for $\\log N(\\HI)\\ge 20.8$ than below this limit (19\\% and 7\\% respectively). The overall H$_2$ detection rate in $\\log N(\\HI)\\ge20$ DLAs is found to be about 16\\% (10\\% considering only $\\log f>-4.5$ detections) after correction for a slight bias towards large $N(\\HI)$. There is a strong preference for H$_2$-bearing DLAs to have significant depletion factors, [X/Fe$]>0.4$. In addition, all H$_2$-bearing DLAs have column densities of iron into dust grains larger than $\\log N($Fe$)_{\\rm dust}\\sim 14.7$, and about 40\\% of the DLAs above this limit have detected H$_2$ lines with $\\log f>-4.5$. This demonstrates the importance of dust in governing the detectability of H$_2$ in DLAs. Our extended sample supports neither the redshift evolution of the detection fraction of H$_2$-bearing DLAs nor that of the molecular fraction in systems with H$_2$ detections over the redshift range $1.8<\\zabs\\le 3$. }% {% }% ", "introduction": "Damped Lyman-$\\alpha$ systems (DLAs) were discovered in the seventies \\citep[e.g., ][]{Lowrance72, Beaver72, Carswell75, Wright79} and identified afterwards as redshifted damping absorptions from large column densities of neutral atomic hydrogen \\citep[][]{Smith79}. Numerous DLAs, with $N(\\HI)\\ge 2\\times 10^{20}$\\,atoms~cm$^{-2}$, have been discovered through large dedicated surveys \\citep[e.g., ][]{Wolfe86} and more recently thanks to the huge number of quasar spectra available from the Sloan Digital Sky Survey \\citep{Prochaska05}. Because DLAs contain most of the neutral hydrogen available for star formation in the Universe \\citep{Wolfe86, Lanzetta91} and are associated with numerous metal absorption lines, they probably arise in the interstellar medium of protogalaxies, progenitors of present-day galaxies \\citep[see, e.g., ][]{Wolfe00, Haehnelt00, Wolfe05}. Our understanding of DLAs is mainly based on the study of low-ionisation metal absorptions \\citep[e.g., ][]{Prochaska02} but also high-ionisation species \\citep{Lu96, Wolfe00, Fox07a, Fox07b} and, in a few cases, molecular absorptions \\citep[e.g., ][]{Ledoux03}. The latter are not conspicuous however in contrast to what is seen in the Galaxy and, for a long time, only the DLA towards Q\\,0528$-$2505 was known to contain H$_2$ molecules \\citep{Levshakov85}. H$_2$-bearing DLAs are nevertheless crucial to understand the nature of DLAs because molecular hydrogen is an important species to derive the physical conditions in the gas \\citep[see, e.g., ][]{Tumlinson02, Reimers03, Hirashita05, Srianand05, Cui05, Noterdaeme07}. The first systematic search for molecular hydrogen in high-redshift ($\\zabs >1.8$) DLAs was carried out using the Ultraviolet and Visual Echelle Spectrograph (UVES) at the Very Large Telescope (VLT) \\citep{Ledoux03}. It consisted of a sample of 33 DLAs with H$_2$ detected in eight of them. Molecular fractions were found to lie in the range $-3.5<\\log f<-1$ with $f=2N($H$_2)/(2N($H$_2)+N(\\HI))$. Upper limits of typically $N($H$_2)\\sim 2\\times 10^{14}$~cm$^{-2}$ (corresponding to $\\log f<-5$) were measured in the other systems. More recently, we noted a correlation between the presence of molecular hydrogen and the metallicity of high-redshift DLAs \\citep{Petitjean06}. High molecular fractions ($\\log f>-4$) were found in about 40\\% of the high-metallicity DLAs ([X/H$]\\ge 1/20^{\\rm th}$ solar) whilst only $\\sim 5$\\% of the [X/H$]<-1.3$ DLAs have $\\log f>-4$. Other papers by our group focused on specific detections. We presented the analysis of three systems with low molecular fractions, i.e., $\\log f<-4$, one of them having a low metallicity \\citep{Noterdaeme07lf}, and the H$_2$-bearing DLA with, to date, the highest redshift, at $\\zabs =4.224$ towards Q\\,1441$+$2737 \\citep{Ledoux06b}. We note that a possible detection of H$_2$ in a DLA towards a Gamma-ray Burst (GRB) afterglow has been reported recently \\citep{Fynbo06}. However, the origin of DLAs at the GRB host-galaxy redshift is very likely to be different from those observed in QSO spectra \\citep[e.g., ][]{Jakobsson06, Prochaska07a}. We present here the whole sample of UVES high-redshift QSO-DLAs for which the wavelength range where H$_2$ lines are redshifted is covered by the available spectra. This sample is more than twice as large as in our previous study \\citep{Ledoux03}. We present the observations and the UVES DLA sample in Sect.~2 and provide comments on individual absorbers in Sect.~3. We discuss the overall population in Sect.~4 and results in Sects.~\\ref{mol_f} to \\ref{z}. We conclude in Sect.~\\ref{conclusion}. ", "conclusions": "} We present results of the largest survey of molecular hydrogen in high-redshift ($1.8<\\zabs \\le 4.2$) DLAs compiled to date using high signal-to-noise ratio, high spectral-resolution VLT-UVES data. We analyse data for 77 DLAs/strong sub-DLAs with $N(\\HI)\\ge 10^{20}$~cm$^{-2}$, a dataset more than twice as large as that studied previously by \\citet{Ledoux03}. From the thirteen high-redshift H$_2$-bearing DLAs known to date, nine have been discovered by our group. Due to the superb quality of the Ultraviolet and Visual Echelle Spectrograph, we are able to detect unambiguously the H$_2$ absorption features, measure accurate column densities, and in the cases of non-detections derive stringent upper limits. A double-sided Kolmogorov-Smirnov test shows that the ten H$_2$-bearing systems with $\\log f>-4.5$ (which is our conservative completeness limit) have \\HI\\ column densities that are compatible with those of the overall DLA population. This may be due however to small number statistics. There is evidence (see Sect.~\\ref{overall_pop}) that the probability of finding large molecular fractions is higher in DLAs with large $N(\\HI)$, as observed in the LMC. About $7$\\% of the systems with $\\log N(\\HI)<20.8$ have $\\log f>-4.5$ while $\\sim 19$\\% of the systems with larger \\HI\\ column densities have similar molecular fractions. There is no sharp transition in the molecular fraction of DLAs at any of the measured total hydrogen ($\\HI+$H$_2$) column density. It is surprising to see that most of the DLAs have very low molecular fractions, i.e., $\\log f \\la -6$. This is much smaller than observed along lines of sight in the Galactic disk \\citep{Savage77}. However, \\citet{Wakker06} measured similar molecular fractions towards high-latitude Galactic lines of sight. We confirm that a good criterion to find H$_2$-bearing DLAs is to select high-metallicity systems. Indeed, 35\\% of the systems with [X/H$]\\ge -1.3$ have $\\log f>-4.5$ whilst this is the case for only 4\\% of those with metallicities lower than that. Since there is a correlation between metallicity and depletion factor, the latter being defined as [X/Fe] with X$=$Zn, S or Si, H$_2$ is found in DLAs also having the highest depletion factors. Therefore, clouds with large molecular fractions are expected to be dusty and clumpy \\citep[e.g.,][]{Hirashita03}. They can be missed because of the associated extinction and/or because of their small cross-section. The presence of H$_2$ is closely related to the dust column density. Indeed, all detections pertain to systems where the column density of iron into dust grains is larger than $5\\times 10^{14}$~cm$^{-2}$, and about 40\\% of these systems have detectable H$_2$. This shows that the presence of dust is an important ingredient in the formation of H$_2$ in DLAs. The low molecular fractions measured for most of the DLAs are probably a consequence of low abundances of metals and dust. We show that the probability of finding H$_2$ increases with increasing velocity width of low-ionisation metal line profiles. The correlation between velocity width and metallicity \\citep{Ledoux06a}, if interpreted as a mass-metallicity relation, provides a natural explanation in which H$_2$-bearing DLA systems are preferably associated with massive objects where star-formation is enhanced. There is no evidence of systematic outflows in H$_2$-bearing DLAs neither from the H$_2$ profiles nor from the study of the associated high-ionisation phase. From the comparison between H$_2$-bearing systems and the overall UVES sample, we show that there is no evolution with redshift of the fraction of H$_2$-bearing DLAs nor of the molecular fraction in systems with detected H$_2$ over the range $1.8<\\zabs\\le 3$. This, compared to the large amounts of H$_2$ observed in the local Universe, suggests that a significant increase of the molecular fraction in DLAs could take place at redshifts $\\zabs\\le 1.8$. Ultraviolet observations from space are therefore needed to observe the H$_2$ Lyman and Werner bands in low and intermediate redshift DLAs. Increasing the sample of H$_2$-bearing DLAs is required to cover a large range in column density and derive the H$_2$ frequency distribution, $f(N($H$_2),${\\sl X}) -- where {\\sl X} is the absorption distance -- in a way similar to studies completed for \\HI\\ \\citep{Lanzetta91}. This would assess whether the steep slope observed in $f(N(\\HI),${\\sl X}) at large column densities \\citep[e.g., ][]{Prochaska05} is due to the conversion \\HI$\\rightarrow$H$_2$ \\citep{Schaye01} rather than to a magnitude-dependent bias from dust obscuration of the background quasars \\citep[e.g., ][]{Boisse98, Vladilo05, Smette05}. Both effects can explain the apparent lack of DLAs of both high metallicity and large \\HI\\ column density, as observed in all DLA samples. While the exact nature of DLAs is still an open debate, H$_2$-bearing DLAs provide interesting probes of the physical conditions close to star-forming regions. About a decade ago, only one H$_2$-bearing DLA was known. With the criteria revealed by the present survey, it will be possible to select these systems more efficiently and their numbers should increase rapidly. This opens up the exciting prospect of the detailed study of the ISM in distant galaxies. \\acknowledgement{We thank an anonymous referee and the language editor for useful comments that improved the paper. PN is supported by a PhD studentship from ESO. PPJ and RS gratefully acknowledge support from the Indo-French Centre for the Promotion of Advanced Research (Centre Franco-Indien pour la Promotion de la Recherche Avanc\\'ee) under contract No. 3004-3. }" }, "0801/0801.3966_arXiv.txt": { "abstract": "We present some results obtained from the synthesis of Stokes profiles in small-scale flux tubes with propagating MHD waves. To that aim, realistic flux tubes showing internal structure have been excited with 5 min period drivers, allowing non-linear waves to propagate inside the magnetic structure. The observational signatures of these waves in Stokes profiles of several spectral lines that are commonly used in spectropolarimetric measurements are discussed. ", "introduction": "The observed wave behaviour in network and facular solar regions, and the role of waves to connect photospheric and chromospheric layers, have drawn a considerable discussion in the recent literature, since it has been proposed that thin magnetic flux tubes in these regions can act as wave guides, supplying energy to the upper layers, in particular for waves with periodicity in the 5-min range \\citep{DePontieu+etal2004}. However, observational studies of the relation between the photospheric and chromospheric signals do not provide a unique picture. It is not clear whether the oscillations remain coherent through the atmosphere \\citep{Lites+Rutten+Kalkofen1993, Krijer+etal2001, Judge+etal2001, Centeno+etal2006b}. There are some hints that the coupling between the photosphere and the chromosphere is within vertical channels \\citep{Judge+etal2001, Centeno+etal2006b}, however, other results point toward inclined wave propagation \\citep{Bloomfield+etal2006, DePontieu+etal2003}. In particular, all these studies point out the important role of the magnetic canopy and the $c_S=v_A$ transformation layer. The dominance of long-period oscillations in the chromosphere of these regions, also remains to be explained. Here we perform numerical simulations of magneto-acoustic waves in small-scale flux tubes, with a subsequent Stokes diagnostics. We address the questions of coherence of oscillations through the atmosphere, wave behaviour at the canopies and the change of the wave period with height. ", "conclusions": "In this paper we have studied the wave behaviour in small-scale flux tubes by means of non-linear numerical simulations and Stokes diagnostics. The following items summarize our conclusions. (i) If the driver that excites oscillations has a vertical component, velocity oscillations with an amplitude of some 150 \\hbox{m$\\;$s$^{-1}$} and Stokes $V$ amplitude oscillations of 10$^{-3}$, in units of the continuum intensity, may be detected in low spatial resolution observations. The oscillations are coherent through the whole atmosphere (with the corresponding time delay due the upward propagation). (ii) If the driver that excites oscillations is purely horizontal, no variations would be detected in observations with reduced spatial resolution, since the vertical velocity variations are produced by the surface mode and are in antiphase on both sides of the tube. (iii) The LOS velocity shows a weak non-linear behaviour in canopy regions, already at heights of 300--350 km in the photosphere. (iv) It is important to take into account the radiative losses of oscillations. They can produce a decrease of the effective cut-off frequency, making possible the propagation of the 5-min oscillations in the chromosphere in vertical small-scale magnetic structures." }, "0801/0801.1708_arXiv.txt": { "abstract": "We present observations of the dwarf galaxies Draco and Ursa Minor, the local group galaxies M32 and M33, and the globular cluster M15 conducted with the Whipple 10m gamma-ray telescope to search for the gamma-ray signature of self-annihilating weakly interacting massive particles (WIMPs) which may constitute astrophysical dark matter (DM). We review the motivations for selecting these sources based on their unique astrophysical environments and report the results of the data analysis which produced upper limits on excess rate of gamma rays for each source. We consider models for the DM distribution in each source based on the available observational constraints and discuss possible scenarios for the enhancement of the gamma-ray luminosity. Limits on the thermally averaged product of the total self-annihilation cross section and velocity of the WIMP, $\\left<{\\sigma}v\\right>$, are derived using conservative estimates for the magnitude of the astrophysical contribution to the gamma-ray flux. Although these limits do not constrain predictions from the currently favored theoretical models of supersymmetry (SUSY), future observations with VERITAS will probe a larger region of the WIMP parameter phase space, $\\left<{\\sigma}v\\right>$ and WIMP particle mass ($m_\\chi$). ", "introduction": "The existence of dark matter (DM) is supported by a variety of observational data including measurements of the cosmic microwave background \\citep{spergel2007}, the large-scale distribution of galaxies \\citep{tegmark2004}, and gravitational lensing \\citep{clowe2006}. In the $\\Lambda$CDM cosmological model that is currently favored by these data, DM comprises approximately $\\sim$26\\% of the total energy density of the universe \\citep{spergel2007}. However, the nature of the particles that constitute DM remains unknown. A popular DM candidate is weakly interacting massive particles (WIMPs), which existed in thermal equilibrium during the early universe and later decoupled as the universe expanded. Since the time of decoupling, the WIMPs have remained non-relativistic, behaving as a collisionless fluid on all but perhaps the shortest spatial scales. In order to reproduce the observed relic density of DM, this hypothetical particle would need to have a cross section on the scale of weak interactions. A stable particle with these properties, the lightest neutralino $\\chi$, can be accommodated in theories of supersymmetry (SUSY). The mass of the neutralino is constrained to be $\\gtrsim$ 6 GeV by CMB measurements and accelerator searches \\citep{bottino2003} and $\\lesssim$ 100 TeV by the unitarity limit on the thermal relic \\citep{griest1990}. In the conventional SUSY scenarios, the neutralino is a Majorana particle which can efficiently self-annihilate in astrophysical environments with high DM density producing secondary particles including high-energy gamma rays. The former and current generation of Air-Cherenkov Telescopes (ACTs), including Whipple, HEGRA, CANGAROO-III, VERITAS, H.E.S.S., and MAGIC are sensitive in the gamma-ray energy range from below 100 GeV to above 10 TeV and can therefore make a substantial contribution to the search for the signatures of DM self-annihilation. Recently several ACTs have detected gamma rays from the Galactic Center (G.C.) \\citep{albert2006,aharonian2004,kosack2004}. Although a more traditional astrophysical origin for this signal is currently favored \\citep{atoyan2004,aharonian2005}, DM self-annihilation has been proposed as a possible explanation for these observations \\citep{horns2005}. We present observations taken with the Whipple 10m telescope of five astrophysical sources with the purpose of detecting the signature of DM self-annihilation. Section \\ref{sourceReviewSection} summarizes the motivations for selecting each source. In Section \\ref{dmFluxSection} we discuss the signature of DM self-annihilation into gamma rays and its dependence on the source astrophysics and the particle physics properties of the WIMP. In Section \\ref{dataSection} we review the atmospheric Cherenkov technique and the methods used to analyze the data. Results of the data analysis are described in Section \\ref{resultsSection}. Models for the DM distribution and scenarios for the enhancement of the gamma-ray flux are presented in Section \\ref{interpretationSection}. We conclude in Section \\ref{susyLimitsSection} by discussing the implications of these observations for the parameter space of allowed SUSY models. ", "conclusions": "We have conducted a search for the gamma-ray signature of neutralino self-annihilation from five sources: the dwarf spheroidal galaxies Draco and Ursa Minor, the globular cluster M15, and the local group galaxies M32 and M33. Each of these sources was chosen as a favorable representative of different astrophysical conditions that could potentially enhance the neutralino density and gamma-ray self-annihilation flux. For a generic MSSM model of the neutralino, the self-annihilation flux is only detectable by the Whipple 10m telescope if such an enhancement is significant. A standard analysis of the data revealed no significant excesses, and upper limits on the gamma-ray flux from each source were derived relative to the flux of the Crab Nebula. We have derived limits on $\\left<{\\sigma}v\\right>$ of the neutralino as a function of its mass using models for the DM profile of each source and the generic differential gamma-ray spectrum of the neutralino self-annihilation constructed from two representative channels, $b\\bar{b}$ and $\\tau^{+}\\tau^{-}$. Using the upper limit on the gamma-ray rate measured from Draco and the most conservative estimate of the DM distribution in this source, we obtain 95 \\% C.L. upper limits on $\\left<{\\sigma}v\\right>$ of $<1.9\\times10^{-21}$ cm$^{3}$ s$^{-1}$ and $<1.2\\times10^{-22}$ cm$^{3}$ s$^{-1}$ for a neutralino of mass 1 TeV annihilating exclusively through the $b\\bar{b}$ and $\\tau^{+}\\tau^{-}$ channels, respectively. We have considered potential enhancements to the DM self-annihilation flux including the effects of DM substructure and the adiabatic compression of the DM halo due to baryonic infall. These scenarios could enhance the annihilation flux by as much as $10^{4}$ and possibly higher. However the uncertainties of these estimates are large due to the poorly understood dynamics in the cores of the objects as well as their potentially complex merging histories. If one assumes that DM is composed of neutralinos with $\\left<{\\sigma}v\\right> \\simeq 3 \\times$ 10$^{-26}$ cm$^{3}$ s$^{-1}$, then some extreme enhancement scenarios for M32 and M33 may be ruled out. The current generation of ACTs such as VERITAS, H.E.S.S., MAGIC, and CANGAROO-III have the potential to improve significantly the sensitivity of these measurements and thus probe a larger region of the MSSM parameter space. For a source with a Crab Nebula-like spectrum, VERITAS has a flux sensitivity $\\sim$10 times better than the Whipple 10m telescope and a peak detection rate near 150 GeV. The lower energy threshold of VERITAS will allow it to be sensitive to the low to intermediate neutralino mass range of 100 GeV--1 TeV. With the improved sensitivity of these instruments, conventional astrophysical backgrounds may become significant. The sensitivity of observations of the G.C. to DM annihilations is already limited by the presence of such backgrounds. Observations of extragalactic sources such as those discussed in this work have the potential to avoid this limitation. Furthermore, the identification of the unique spectral signature of DM self-annihilations in two or more of these sources would effectively rule out a traditional astrophysical process. Next-generation ACT instruments such as the currently planned Cherenkov Telescope Array (CTA) and Advanced Gamma Ray Imaging System (AGIS) will potentially be 10$^{2}$--10$^{3}$ times as sensitive as the Whipple 10m telescope and could perform dedicated deep observations with 10--10$^{2}$ times longer exposure than the observations presented in this work. This may allow the exclusion of MSSM models with even the most conservative assumptions for the DM distribution in the sources with the lowest anticipated astrophysical backgrounds such as the dwarf spheroidal galaxies." }, "0801/0801.2017_arXiv.txt": { "abstract": "We estimate the amplitude of perturbation in dark energy at different length scales for a quintessence model with an exponential potential. It is shown that on length scales much smaller than hubble radius, perturbation in dark energy is negligible in comparison to that in in dark matter. However, on scales comparable to the hubble radius ($\\lambda_{p}>1000\\mathrm{Mpc}$) the perturbation in dark energy in general cannot be neglected. As compared to the $\\Lambda$CDM model, large scale matter power spectrum is suppressed in a generic quintessence dark energy model. We show that on scales $\\lambda_{p} < 1000\\mathrm{Mpc}$, this suppression is primarily due to different background evolution compared to $\\Lambda$CDM model. However, on much larger scales perturbation in dark energy can effect matter power spectrum significantly. Hence this analysis can act as a discriminator between $\\Lambda$CDM model and other generic dark energy models with $w_{de} \\neq -1$. ", "introduction": "Cosmological observations suggest that about $70\\%$ content of our universe is made of a form of matter which drives the accelerated expansion of the universe\\cite{obs_proof}. These observations include Supernova type Ia observations \\cite{nova_data1}, observations of Cosmic Microwave Background \\cite{boomerang,wmap_params,2003Sci...299.1532B} and large scale structure \\cite{2df, sdss, 2004PhRvD..69j3501T}. The accelerated expansion of the universe can of course be explained by introducing a cosmological constant $\\Lambda$ in the Einstein's equation \\cite{ccprob_wein,review3}. However, the cosmological constant model is plagued by the fine turning problem \\cite{ccprob_wein}. This has motivated the study of dark energy models to explain the current accelerated expansion of the universe. The simplest model as an alternative to cosmological constant model is to assume that this accelerated expansion is driven by a canonical scalar field with a potential $V(\\phi)$. This class of dark energy models are known as quintessence models and the scalar field is known as a quintessence field. Various quintessence models have been studied in literature \\cite{quint1,ferreira,liddle}. There exists another class of string theory inspired scalar field dark energy models known as tachyon models \\cite{tachyon1,2003PhRvD..67f3504B}. Models of dark energy which allow $w<-1$ are known as phantom models \\cite{2002PhLB..545...23C}. Phantom type dark energy can also be realized in a scalar tensor theory of gravitation. (See for example Ref.\\cite{STG}.) Other scalar field models include k-essence field \\cite{2001PhRvD..63j3510A}, branes \\cite{brane1}, Chaplygin gas model and its generalizations \\cite{chaply}. There are also some phenomenological models \\cite{water}, field theoretical and reorganization group based models (see e.g. \\cite{tp173}), models that unify dark matter and dark energy \\cite{unified_dedm1}, holographic dark energy models \\cite{HGDE} and many others like those based on horizon thermodynamics (e.g. see \\cite{2005astro.ph..5133S}). For reviews of dark energy models see for instance Ref.\\cite{DEreview} and for constraining parameters using observations see Ref.\\cite{param_fit}. Homogeneous dark energy distribution leads to accelerated expansion of the Universe which, in turn, governs the luminosity distance and angular diameter distance. The rate of expansion also influences growth of density perturbations in the universe. This is evident from the abundance of rich clusters of galaxies and their evolution and the Integrated Sachs Wolfe effect \\cite{isw0}. In this paper we present a set of argument which lead to the conclusion that inhomogeneous dark matter with homogeneous dark energy at all length scales is inconsistent with Einstein's equation if $p_{de} \\neq -\\rho_{de}$. We further analyze how dark matter power spectrum is influenced by perturbation in dark energy with an evolving equation of state parameter. Dark energy perturbations have been extensively studied in the linear approximation \\cite{weller_lewis,bean_dore,depert}. It was shown in Ref.\\cite{weller_lewis} that dark energy perturbations affect the low $l$ quadrapole in the CMB angular power spectrum through the ISW effect. This analysis was done for a constant equation of state parameter. For models with $w>-1$ this effect is enhanced while for phantom like models it is suppressed. In these models dark matter perturbations and dark energy perturbations are anti-correlated for large effective sound speeds. This anti-correlation is a gauge dependent effect \\cite{bean_dore}. Detailed studies in dark energy perturbations also include \\cite{chpgas_pert,sph_coll}. Clustering of dark energy within matter over density and voids were studied by Mota et al\\cite{mota}. In this paper we study the evolution of perturbation in dark energy in a quintessence model which results in an evolving equation of state parameter different from that considered in Refs.\\cite{amendola1, weller_lewis}. We use a specific model of scalar field dark energy with an exponential potential. We find that although for scales much smaller than hubble radius the perturbation in dark energy is small, for scales $\\geq H^{-1}$ the perturbation in dark energy can be comparable to that in matter. Hence, although on small scales ($<< H^{-1}$) the dark energy can be treated as homogeneous, one has to take into account the perturbation in dark energy over scales $\\sim H^{-1}$ if $w_{de} \\neq -1$. Clearly, in the specific case of $w_{de} = -1$ the dark energy is homogeneous at all scales. We choose to work in the longitudinal gauge as in that case we can directly relate the metric perturbation $\\Phi$ to the gravitational potential perturbation. For a specific model, we investigate how quintessence dark energy influences matter power spectrum. We show that on scales $\\lambda_{p} < 1000\\mathrm{Mpc}$, the matter power spectrum is not significantly affected whether or not we include fluctuations in the quintessence field in perturbation equations. However, on much larger scales, including or excluding fluctuations in quintessence field does result in significant changes in the matter power spectrum. This paper is organized as follows. In Section \\ref{sec::inhomogeneous DE} we discuss the background cosmology for matter and the scalar field system and describe the cosmological perturbation equation in longitudinal gauge for this system. In Section \\ref{sec::Numerical Solutions} we obtain the numerical solution of the perturbation equation to determine the ratio of the perturbations in dark energy to the perturbations in matter at various length scales. Section \\ref{sec::Conclusions} summarizes the results. ", "conclusions": "\\label{sec::Conclusions} In this paper, we have investigated the perturbations in dark energy. This is motivated by the fact that the assumption that the distribution of dark energy (with $w \\neq -1$) is homogeneous at all length scales is inconsistent with the observational fact that dark matter is distributed inhomogeneously. On length scales comparable to or greater than the Hubble radius ($\\lambda_{p}> 1000\\mathrm{Mpc}$), the perturbation in dark energy can become comparable to perturbation in matter if $w_{de} \\neq -1$. The model parameters we have chosen correspond to $w \\approx -0.8 $ and $w \\approx -0.9$, which are within the range allowed by Supernova observations and WMAP5 observations. Given this range, the evolution of perturbations differs significantly. For scales $\\lambda_{p}< 1000\\mathrm{Mpc}$, the perturbation in dark energy $\\delta_\\phi$ can be neglected in comparison with the perturbation in matter $\\delta_{m}$ at least in the linear regime. We have demonstrated this using an exponential potential for the quintessence field. This result agrees with those presented in Ref.\\cite{mota} on sub-Hubble scales. We have further demonstrated that quintessence dark energy results in suppression of matter power spectrum relative to $\\Lambda CDM$ model. We found that at length scale of $\\lambda_{p}= 1000\\mathrm{Mpc}$ and for the value of the parameter $\\lambda = 1$, the matter power spectrum is suppressed by about $4\\%$ compared to its value in the $\\Lambda$CDM model for the same set of initial condition. However, at $\\lambda_{p}= 10^{5}\\mathrm{Mpc}$, matter power spectrum is suppressed by about $15\\%$ compared to its value in the $\\Lambda$CDM model. We have demonstrated that on scales $\\lambda_{p}< 1000\\mathrm{Mpc}$ this suppression is primarily due to different background evolution relative to $\\Lambda$CDM model. The resultant matter power spectrum is nearly invariant even if we assume that quintessence field is homogeneous on these scales. However, on much larger scale $\\lambda_{p}> 1000\\mathrm{Mpc}$, including or excluding fluctuation in the quintessence field results in significant changes in the matter power spectrum. All these results emphasize that dark energy can indeed be treated as nearly homogeneous on scales $\\lambda_{p}< 1000\\mathrm{Mpc}$. However, on much larger scales ($\\lambda_{p} > 1000\\mathrm{Mpc}$), if the equation of state parameter deviates from -1, then perturbations in dark energy does influences matter power spectrum significantly. If a definitive detection of perturbations in dark energy is made, it will certainly rule out the cosmological constant at least as the sole candidate of dark energy." }, "0801/0801.0154_arXiv.txt": { "abstract": "% We describe very accurate imaging of radio spectral index for the inner jets in three FRI radio galaxies. Where the jets first brighten, there is a remarkably small dispersion around a spectral index of 0.62. This is also the region where bright X-ray emission is detected. Further from the nucleus, the spectral index flattens slightly to 0.50 - 0.55 and X-ray emission, although still detectable, is fainter relative to the radio. The brightest X-ray emission from the jets is therefore not associated with the flattest radio spectra, but rather with some particle-acceleration process whose characteristic energy index is 2.24.The change in spectral index occurs roughly where our relativistic jet models require rapid deceleration. Flatter-spectrum edges can be seen where the jets are isolated from significant surrounding diffuse emission and we suggest that these are associated with shear. ", "introduction": "\\label{radio} The detection of X-ray synchrotron emission on kiloparsec scales in the bases of low-luminosity (FR\\,I) radio jets (e.g.\\ \\citealt{Hard02,Hard05,ngc315xray}) requires distributed particle acceleration, consistent with the failure of adiabatic models to reproduce the observed brightness profiles at radio wavelengths \\citep{LB04}. The acceleration mechanism is not understood. Here, we summarize the results of recent work on this problem, using accurate radio spectral mapping (this section) and detailed radio -- X-ray comparisons (Section~\\ref{xray}). In the course of our jet-modelling programme \\citep[and references therein]{L07}, we have accumulated high-resolution, multi-frequency radio images of the bases of FR\\,I jets and have used these to derive accurate spectral-index distributions. We have studied NGC\\,315 \\citep{ngc315ls} and 3C\\,296 \\citep{3c296}; a third example, the nearby radio galaxy 3C\\,31 (z = 0.0169; Laing et al., in preparation), is shown in Fig.~\\ref{fig:3c31alpha}. There is no evidence for any significant deviation from power-law spectra in 3C\\,31 within 70\\,arcsec of the nucleus. Out to $\\approx$7\\,arcsec in both jets, the spectral index at 1.5-arcsec resolution is slightly steeper ($\\langle \\alpha \\rangle = 0.62$)\\footnote{$S(\\nu) \\propto \\nu^{-\\alpha}$} than the average for the inner jets. From 7 -- 50\\,arcsec in both jets, the mean spectral index is in the range 0.55 -- 0.57. Further from the nucleus, there is a gradual spectral steepening. There are also slight, but significant variations in spectral index across both jets within $\\approx$30\\,arcsec of the nucleus in the sense that their West edges tend to have flatter spectra ($\\langle \\alpha \\rangle = $ 0.52 -- 0.54; Fig.~\\ref{fig:3c31alpha}c). There is a clear spectral separation between the South jet and the surrounding emission, matching the separation of these regions defined by the sharpest brightness gradients (Figs~\\ref{fig:3c31alpha}a and b). This is particularly clear where the jet first enters the diffuse emission. The spectral identity of the jet is evidently maintained even after it bends abruptly about 2\\,arcmin South of the nucleus and remains until it terminates in a region of high brightness gradient. The outer South jet in 3C\\,31 is therefore a clear example of the type of spectral structure noted in other FR\\,I sources \\citep{KSR97,3c296} wherein a flatter-spectrum `jet' with a distinct spectral identity is superposed on a `sheath' of steeper-spectrum emission. \\begin{figure} \\plotone{3c31_alphabw.eps} \\caption{Radio images of 3C\\,31. (a) Sobel-filtered, mean L-band image (normalized by total intensity) at a resolution of 5.5\\,arcsec. (b) and (c) Spectral index, $\\alpha$ from weighted least-squares, power-law fits to the total intensity. (b) 5-frequency fit between 1365 and 4985\\,MHz at a resolution of 5.5\\,arcsec FWHM. (c) 6-frequency fit between 1365 and 8440\\,MHz at a resolution of 1.5\\,arcsec FWHM for the inset area in panel (b).\\label{fig:3c31alpha}} \\end{figure} ", "conclusions": "These observations contribute to the developing picture of spectral variations in FR\\,I sources. Bright X-ray emission is detected close to the nucleus, in the faint, well-collimated jet bases that precede the sudden radio brightening (e.g.\\ Fig.~\\ref{fig:xprofiles}). There is approximate morphological correspondence between features in the radio and X-ray brightness distributions after the former brightens, although there are differences on small scales (e.g.\\ Fig.~\\ref{fig:ngc315xray}b). In contrast, there are no systematic transverse variations in the X-ray/radio ratio within $\\approx$30\\,arcsec of the nucleus in NGC\\,315 (the best resolved case; \\citealt{ngc315xray}). Particle acceleration appears to be distributed throughout the jet volume, rather than being exclusively associated with discrete knots or with the boundary. The ratio of X-ray to radio emission decreases where our kinematic models show that the jets start to decelerate from speeds of 0.8 -- 0.9$c$ (Fig.~\\ref{fig:xprofiles}; \\citealt{LB02,Hard02,CLBC,ngc315xray}). Where the jets first brighten and before they decelerate, there is a remarkably small dispersion around a radio spectral index of $\\alpha = 0.62$ in the three sources we have studied in detail, as well as 3C\\,66B (Fig.~\\ref{fig:3c31alpha}; \\citealt*{HBW,ngc315ls,3c296}). The average is dominated by emission immediately after the point at which the jets first brighten. This is also the region from which X-ray emission is detected from the main jets in all four sources (Fig.~\\ref{fig:xprofiles}; \\citealt{HBW,Hard05}). The spectral index of the fainter emission close to the nucleus in 3C\\,449 \\citep{KSR97}, PKS1333$-$33 \\citep{KBE} and 3C\\,66B \\citep{HBW} appears to be slightly steeper than $\\alpha = 0.62$ although the uncertainties are larger. Further from the nucleus, the spectra flatten slightly to $\\alpha = $ 0.50 -- 0.55, contrary to any naive expectation from models in which electrons are accelerated at the brightening point and suffer synchrotron losses as they propagate. X-ray emission is still detected from these regions, but at a lower level relative to the radio (Fig.~\\ref{fig:xprofiles}). The brightest X-ray emission from the jets is therefore not associated with the flattest radio spectra, but rather with some particle acceleration process whose characteristic energy index is $2\\alpha + 1 = 2.24$. A related result is that an asymptotic low-frequency spectral index of 0.55 is common in FR\\,I jets over larger areas than we consider here \\citep{Young}. Flatter-spectrum edges can be seen where the jets are isolated from significant surrounding diffuse emission, most clearly in NGC\\,315 \\citep{ngc315ls}. Our kinematic models \\citep{LB02,CLBC,3c296} show that all of the jets have substantial transverse velocity gradients and it is plausible that the process that produces the flatter spectrum is associated with high shear \\citep{SO}. In 3C\\,31, the flatter-spectrum regions (Fig.~\\ref{fig:3c31alpha}c) occur predominantly on the outer edges of bends, perhaps consistent with this idea. As well as a smooth steepening of the jet spectrum at larger distances from the nucleus, as would be expected from synchrotron and adiabatic losses affecting a homogeneous electron population, multiple spectral components are observed (Fig.~\\ref{fig:3c31alpha}b). Jets appear to retain their identities even after entering regions of diffuse emission and are clearly identifiable by their flatter spectra. They are usually also separated from the surrounding emission by sharp brightness gradients (Fig.~\\ref{fig:3c31alpha}a). This spine/sheath separation is observed in FR\\,I sources with bridges of emission extending back towards the nucleus (e.g.\\ 3C\\,296; \\citealt{3c296}) as well as tailed sources like 3C\\,31 (Fig.~\\ref{fig:3c31alpha}). Although there is an overall trend for the spectrum of the diffuse emission to steepen towards the nucleus in bridges and away from it in tails \\citep{Parma99}, the variations in individual objects are complex. The termination regions of jets in tailed FR\\,I sources are perhaps best regarded as bubbles which are continually fed with fresh relativistic plasma by the jets and which in turn leak material into the tails. Their spectral steepening would then be governed by a combination of continuous injection, adiabatic, synchrotron and inverse Compton energy losses and escape." }, "0801/0801.0965_arXiv.txt": { "abstract": "As part of a large survey of halo and thick disc stars, we found one halo star, HD 106038, exceptionally overabundant in beryllium. In spite of its low metallicity, [Fe/H] = $-$1.26, the star has log(Be/H) = $-$10.60, which is similar to the solar meteoritic abundance, log(Be/H) = $-$10.58. This abundance is more than ten times higher the abundance of stars with similar metallicity and cannot be explained by models of chemical evolution of the Galaxy that include the standard theory of cosmic-ray spallation. No other halo star exhibiting such a beryllium overabundance is known. In addition, overabundances of Li, Si, Ni, Y, and Ba are also observed. We suggest that all these chemical peculiarities, but the Ba abundance, can be simultaneously explained if the star was formed in the vicinity of a hypernova. ", "introduction": "The single stable isotope of beryllium, $^{9}$Be, is a pure product of cosmic-ray spallation of heavy (mostly CNO) nuclei \\citep*{RFH70}. Analyses of Be abundances in metal poor stars \\citep{MBCP97,BDKR99} have found a relationship between [Fe/H]\\footnote{[A/B] = log [N(A)/N(B)]$_{\\rm \\star}$ - log [N(A)/N(B)]$_{\\rm\\odot}$} and log(Be/H) with slope close to one, and between [O/H] and log(Be/H) with slope between 1 and 1.5, depending on the oxygen indicator used. Independently of the behaviour of the [O/Fe] ratio at lower metallicities, these results suggest a primary production of Be in the early Galaxy \\citep{Ki01}. As a primary element, and assuming cosmic-rays to be globally transported across the early Galaxy, Be may show a smaller scatter than the products of stellar nucleosynthesis \\citep{SY01} at a given time, suggesting its potential use as a cosmochronometer. So far, the linear relations appear to be very tight, showing a surprisingly low scatter comparable to the measurement errors. This picture, however, might change with the increase of the samples analysed, as hinted by the results of \\citet{BN06}. Nevertheless, turn-off stars of the globular clusters NGC 6397 and NGC 6752 were found \\citep{Pas04,Pas07} to have the same Be abundance of field stars of the same metallicity. This strongly support the production of Be to be a global process. Ages derived from these abundances, in a comparison with a model of the evolution of Be with time \\citep{Va02}, show an excellent agreement with ages derived from theoretical isochrones, supporting the use of Be as a cosmochronometer. Moreover, \\citet{Pas05} showed that Be abundances could be used to study the differences in the time scales of star formation in the halo and the thick disc of the Galaxy. In this letter, we report the discovery of an extremely Be enriched halo star, HD 106038, with an abundance 1.2 dex higher than stars of similar metallicity. This unique star deviates considerably from the observed relations of Be with Fe and O. It was identified during the analysis of a large sample containing near to one hundred halo and thick disc stars (Smiljanic et al. 2008, in preparation). Neither the standard scenario for Be production, involving spallation of cosmic-rays on nuclei of the interstellar medium \\citep{Va02}, nor the superbubbles (SBs) scenario \\citep{Par00} seem to be able to produce such Be enriched objects. The SBs model predict a scatter in the Be abundance \\citep{PD00} that may explain the stars found by \\citet{BDKR99} and \\citet{BN06} which have similar atmospheric parameters but Be abundances differing by $\\sim$ 0.5 dex. The very high Be abundance in HD 106038, however, would require an extremely poor mixing of the SNe ejecta with the ISM which seems to be difficult to justify \\citep{Par00}. \\begin{figure} \\begin{centering} \\includegraphics[width=6.5cm]{smiljanic_fig01.eps} \\caption{Comparison between the spectra of HD 106038 (solid line) and of HIP 7459 (dashed line), a star with close atmospheric parameters and similar metallicity, in the Be region. The dominating element of the nearby blended features are also indicated. The V and Ti features have the same strength in the two stars while some difference in the Zr line is noted.} \\label{fig:be} \\end{centering} \\end{figure} To the best of our knowledge, there is only one other case of extremely Be enhanced star in the literature.\\footnote{We exclude from the discussion the chemically peculiar A or F stars with enhanced Be lines. The peculiar abundances of these stars are thought to be caused by effects of diffusion. As shown by \\citet*{Ric02}, these effects do not result in overabundances in stars with similar temperature and metallicity as HD 106038.} The star J37 of the open cluster NGC 6633 was found by \\citet{As05} to have log (Be/H) = $-$9.0 $\\pm$ 0.5. The chemical peculiarities of star J37 might be best explained by the accretion of rocky material similar to chondritic meteorites \\citep{As05}. As we shall see below, the accretion of such a material is unlikely for our population II star. ", "conclusions": "Since the standard scenario for cosmic-ray spallation does not explain the enhancement of Be in HD 106038, a peculiar and/or rare event may be related to its formation. A combination of two or more rare events to produce the observed features is unlikely, we therefore concentrate on single events. To reproduce the very particular chemical pattern of HD 106038, a nucleosynthetic site must be able to overproduce Si and Ni without overproducing other $\\alpha$ and iron group elements. Elements in normal halo stars with the same metallicity as HD106038 come mostly from SNe II. It is therefore unlikely that the same SNe II may produce the observed enhanced [Si/Fe] and [Ni/Fe] ratios. Moreover, it has about 16 times more Be than what models involving SN II predict for its metallicity \\citep{Va02}. SNe Ia produce large amounts of Fe, Ni, and other iron group elements but are not expected to produce large amounts of O and Si \\citep{Iw99}. Therefore, in case of a contribution from SNe Ia, ratios such as [O/Fe] and [Si/Fe] would fall below normal halo stars, contrary to what is observed. Moreover, SNe Ia are expected to produce about one order of magnitude less spallation products than SN II \\citep{Fie02}. All these features, however, may possibly be found in the ejecta of a hypernova (HNe), which can be enriched in both intermediate mass elements (as S and Si) as in iron group elements \\citep{Nak01,Pod02}. Moreover, HNe can produce large amounts of Be and Li by spallation \\citep{Fie02,Nak04}. Therefore, we suggest that the material which formed this star was probably contaminated by the nucleosynthetic products of a HNe. \\subsection{Hypernova}\\label{sec:hyp} Hypernovae are core-collapse SNe (usually of type Ic) with exceptionally large kinetic energy production, resulting in spectra dominated by very broad absorption line blends \\citep{Maz00}. The energy released in the explosion can be one order of magnitude larger than that of normal core-collapse SNe \\citep{Iw03}. Some hypernovae, typically the most massive and energetic events, are linked to Gamma-Ray bursts \\citep{Iw98}. \\citet{Fie02} and \\citet{Nak04} calculated the yields of spallation products resulting from HNe explosions. While \\citet{Nak04} calculate the energy distribution of the ejecta with a hydrodynamic code and solve the cosmic-ray transfer equation, \\citet{Fie02} use an empirical formula for the energy distribution and do not solve the transfer equation but adopt an approximation to have the mass fraction of the ejecta that produces the spallation products. \\citet{Nak04} claim the simplifications adopted by \\citet{Fie02} to overestimate the yields by a factor $\\sim$ 3. The yield of Be per HNe can be one or two orders of magnitude larger than the one per SNe II \\citep{Fie02,Nak04}. However, as a rare event, they are not major contributors of Be in the Galaxy. \\citet{Fie02} predict $^{7}$Li/$^{9}$Be $\\sim$ 8.6 and Be/O $\\sim$ $5.6 \\times 10^{-7}$ (both ratios by number). The calculations by \\citet{Nak04} predict $^{7}$Li/$^{9}$Be $\\sim$ 4.2, also by number. Both predictions are close to what is observed in HD 106038; the ratio between the observed excess of $^{7}$Li and $^{9}$Be is $^{7}$Li/$^{9}$Be = 5.6 (while the HNe is expected to have produced all the observed Be abundance, it is responsible only for the excess of Li with respect to the primordial plateau). We cannot estimate the contribution of the possible HNe on the observed oxygen abundance, thus only a lower limit can be placed, Be/O \\textgreater $1.9 \\times 10^{-7}$, given by the assumption that all the observed oxygen has been produced by the HNe. Both models, however, predict much more $^{6}$Li than observed, $^{7}$Li/$^{6}$Li $\\sim$ 1.9 by \\citet{Fie02} and $^{7}$Li/$^{6}$Li $\\sim$ 1.2 by \\cite{Nak04}, by number. The observed ratio between the excess of $^{7}$Li and the $^{6}$Li abundance is $^{7}$Li/$^{6}$Li $\\leq$ 15. We consider this ratio an upper limit since, given its fragility, some $^{6}$Li has probably been destroyed in previous evolutionary phases. The production of Be without a corresponding production of $^{6}$Li would be extremely difficult to understand. Nucleosynthetic calculations by \\citet{Nak01} find the ejecta of HNe to have smaller amounts of C and O and larger amounts of Si, S, and Ar, when compared to normal SNe. \\citet{Nak01} and \\citet{No06} also note larger [(Zn,Co)/Fe] and smaller [(Mn,Cr)/Fe] ratios. An overabundance of Zn, which is not observed, can be avoided with a deeper mas cut\\footnote{The coordinate, in mass, separating the part of the star that is ejected and the one that forms the remnant.}, which would also result in a larger [Ni/Fe] \\citep{No06}. These are in qualitatively agreement with the observations, supporting the HNe hypothesis. The weak s-process in massive stars seems to efficiently produce only elements with a mass number\\footnote{The mass number of Y is A = 89 and of Ba is A $\\sim$ 136.} up to 90 \\citep{RH00}. The Y overabundance may require an enhanced flux of neutrons, which would also contribute to the production of Ni. The Ba overabundance, however, is more difficult to understand. It is not clear whether this same mechanism would result in the overproduction of Ba. Moreover, a significant amount of Ba is expected to be produced only by the main s-process in AGBs or by the r-process, usually associated to massive stars. Since pollution by AGBs is not possible (see below), the Ba overabundance is likely a product of a massive star. For example, although not expected, \\citet*{Maz92} found Ba to be overabundant by a factor of 5 in the spectrum of SN 1987A. Although Ba might pose a problem for our scenario, we recall that theoretical predictions for the r-process elements in HNe are not available. Therefore, whether this scenario is able to explain the Ba abundance is still an open question. The HNe scenario, at least qualitatively, is able to explain most features observed in HD 106038 within a single peculiar event. More work, however, is still needed to show whether the scenario still holds quantitatively. A detailed comparison with nucleosynthetic predictions of theoretical models is necessary to validate or not the HNe hypothesis. \\subsection{Other scenarios}\\label{sec:oth} In this subsection we present some alternative scenarios for the origin of the Be enhancement in HD 106038, which were discarded for the reasons presented below. A pollution by AGB stars, or any kind of evolved star, can at once be discarded. Although these may explain the s-process elements, and maybe Li, they do not explain the Be overabundance. On the contrary, the material ejected by an AGB or by a massive star, after successive mixing events, would be depleted in Be. The SBs scenario \\citep{Par00} would be able to reproduce the observed Li/Be and Be/O ratios only with an extreme model where particles are accelerated from pure SNe ejecta. The SBs evolution models, however, predict that most material inside of a SB comes from the ISM. Moreover, the remaining chemical peculiarities are not typical of SNe II ejecta and thus can not be explained within the same scenario. Another possibility is the engulfing of a sub-stellar object, a planet or planetesimals debris as in star J37 of the open cluster NGC 6633 \\citep{As05}. This, however, can be excluded with a robust quantitative argument. The accreted material would be confined to the surface convective layer of the star. In a metal poor star this layer is much shallower than in a solar metallicity star. With the equation given by \\citet{Mu01} we estimate the surface convective layer of HD 106038 to have $\\sim$ $4.5 \\times 10^{-3}$ M$_{\\odot}$. The mass of Be in this layer is $\\sim$ $7.7 \\times 10^{-13}$ M$_{\\odot}$ while in a star with normal abundance of Be it would be $\\sim$ $4.8 \\times 10^{-14}$ M$_{\\odot}$. Assuming the accreted material to have a composition similar to chondrites meteorites \\citep{Lod03} a mass of Fe of $5.3 \\times 10^{-6}$ M$_{\\odot}$ would also be accreted with the required mass of Be. However, the total mass of Fe in the convective layer of the star is $\\sim$ $3.3 \\times 10^{-7}$ M$_{\\odot}$. In this scenario all, or almost all, Fe in the convective layer of the star would come from the accreted material. HD 106038 would then originally be a population III metal free star; an extremely unlikely possibility. In addition, we note that the most metal poor star found to host planets has [Fe/H] = $-$0.68 \\citep{Co07}. A longer exposure to EPs would be the natural explanation if HD 106038, for some reason, was younger than halo stars of the same metallicity. Its Be abundance would be a result of the accumulated action of EPs in a cloud where star formation was, somehow, delayed. Its Be abundance higher than the solar photospheric one could be a sign of a solar or younger age, in agreement with the suggestion of \\citet{Pas04} that Be abundances could be used as a cosmic clock. The position of the star in the HR diagram, although not favourable for a good age determination, favours an older age and argues against this hypothesis. If the star originated in or near the Galactic centre, the enhanced star formation and supernovae events could provide an enhanced EPs flux. This flux might also originate from a non-stellar source such as the central black hole. A bulge origin for the star, however, seems unlikely. In particular the abundances of Ni, Y, and Ba are not compatible with the ones of bulge stars \\citep{MR94}. Since it requires another source for the Ni and s-process overabundances, we also discard this hypothesis." }, "0801/0801.3304_arXiv.txt": { "abstract": "{A promising method for the detection of UHE neutrinos is the Lunar Cherenkov technique, which utilises Earth-based radio telescopes to detect the coherent Cherenkov radiation emitted when a UHE neutrino interacts in the outer layers of the Moon. The LUNASKA project aims to overcome the technological limitations of past experiments to utilise the next generation of radio telescopes in the search for these elusive particles. To take advantage of broad-bandwidth data from potentially thousands of antennas requires advances in signal processing technology. Here we describe recent developments in this field and their application in the search for UHE neutrinos, from a preliminary experiment using the first stage of an upgrade to the Australia Telescope Compact Array, to possibilities for fully utilising the completed Square Kilometre Array. We also explore a new real time technique for characterising ionospheric pulse dispersion which specifically measures ionospheric electron content that is line of sight to the moon.} \\begin{document} ", "introduction": "The origin of the most energetic particles observed in nature, the ultra high energy (UHE) cosmic rays (CR), which have energies extending up to at least $2 \\times 10^{20}$ eV, is currently unknown. Finding the origin of these particles will have important astrophysical implications. However, direct detection of UHE neutrinos is very difficult due to their extremely small interaction cross-sections. Instead, they may be detected indirectly via observation of the Askaryan effect \\cite{askaryan62excess} in the lunar regolith. Askaryan first predicted coherent Cherenkov emission in dielectric media at radio and microwave frequencies. Using the Moon as a large volume neutrino detector, coherent radio Cherenkov emission from neutrino-induced cascades in the lunar regolith can be observed with ground based telescopes. This method was first proposed by Dagkesamanskii and Zheleznykh \\cite{dagkesamanskii89radio} and first applied by Hankins, Ekers and O'Sullivan \\cite{hankins96asearch} using the Parkes radio telescope. Coherent Cherenkov radiation is a linearly polarised broadband emission. The spectrum of coherent Cherenkov emission rises approximately linearly with frequency until a peak value is reached. The peak frequency is determined by de-coherence and/or attenuation in the regolith, and can vary between a few hundred MHz and approximately 5 GHz. The dependence of the peak frequency on shower geometry makes the choice of an optimum observation frequency non-trivial \\cite{clancy07lunar}. ", "conclusions": "" }, "0801/0801.3845_arXiv.txt": { "abstract": "The formation of disk galaxies is one of the most outstanding problems in modern astrophysics and cosmology. We review the progress made by numerical simulations carried out on large parallel supercomputers. These simulations model the formation of disk galaxies within the current structure formation paradigm in which the Universe is dominated by a cold dark matter component and a cosmological constant. We discuss how computer simulations have been an essential tool in advancing the field further over the last decade or so. Recent progress stems from a combination of increased resolution and improved treatment of the astrophysical processes modeled in the simulations, such as the phenomenological description of the interstellar medium and of the process of star formation. We argue that high mass and spatial resolution is a necessary condition in order to obtain large disks comparable with observed spiral galaxies avoiding spurious dissipation of angular momentum. A realistic model of the star formation history. gas-to-stars ratio and the morphology of the stellar and gaseous component is instead controlled by the phenomenological description of the non-gravitational energy budget in the galaxy. This includes the energy injection by supernovae explosions as well as by accreting supermassive black holes at scales below the resolution. We continue by showing that simulations of gas collapse within cold dark matter halos including a phenomenological description of supernovae blast-waves allow to obtain stellar disks with nearly exponential surface density profiles as those observed in real disk galaxies, counteracting the tendency of gas collapsing in such halos to form cuspy baryonic profiles. However, the ab-initio formation of a realistic rotationally supported disk galaxy with a pure exponential disk in a fully cosmological simulation is still an open problem. We argue that the suppression of bulge formation is related to the physics of galaxy formation during the merger of the most massive protogalactic lumps at high redshift, where the reionization of the Universe likely plays a key role. A sufficiently high resolution during this early phase of galaxy formation is also crucial to avoid artificial angular momentum loss and spurious bulge formation. Finally, we discuss the role of mergers in disk formation, adiabatic halo contraction during the assembly of the disk, cold flows, thermal instability and other aspects of galaxy formation, focusing on their relevance to the puzzling origin of bulgeless galaxies. ", "introduction": "Galaxies occupy a special place in our quest for understanding the Universe. They are large islands in a nearly empty space and contain most of the ordinary baryonic matter, stars and interstellar gas, that emits radiation and can thus be detected by astronomers$^1$. Galaxies come essentially in two broad categories$^2$, those in which the luminous mass is arranged in a rotating disk of stars and gas, called disk galaxies or spiral galaxies because of the presence of spiral arms of gas and stars (Figure 1), and those in which the luminous mass is distributed in a smooth, featureless spheroidal structure with little or no rotation, also known as elliptical galaxies (Figure 1). Disk galaxies have a radial light distribution $I(r)$ that is well fit by a decaying exponential law$^3$, $I(r) \\sim exp(-r/r_d)$, where $r_d$ is a characteristic scale length ($r_d \\sim 2-4$ kpc for typical spiral galaxies). Indeed many disk galaxies contain also a spheroidal stellar component at their center, the stellar bulge, which has structural properties similar to an elliptical galaxies albeit being much smaller in size$^2$. Both types of galaxies are known to contain dark matter, namely matter that is not traced by radiation. In disk galaxies dark matter clearly dominates over luminous matter by mass, as inferred from their high rotation speeds which requires the gravitational pull of a massive and extended halo of dark matter$^4$. We live in a galaxy of the first kind, the Milky Way. Indeed disk galaxies are ubiquitous in the local Universe, and also at the largest distances and earliest epochs at which the best ground and space-based telescopes have been able to study the morphology of galaxies reliably$^5$. Only the most massive galaxies in the Universe do not posess a disk component, while this becomes progressively more dominant compared to the spheroidal component as the mass of the galaxy decreases (Figure 2). \\begin{figure} \\centering{ \\resizebox{12cm}{!}{\\includegraphics{gdesignlr.ps}}\\\\ \\resizebox{7cm}{!}{\\includegraphics{m87lr.ps}}\\\\ \\caption{Top: a typical disk-dominated galaxy, the nearby spiral galaxy M102 in the Ursa Major constellation, 27 million light years from the Sun (the image was obtained by Chris \\& Dawn Schur from Payson, Arizona at 5150 feet elevation with an amateur telescope). A small spheroidal bulge is visible at the center of the disk. Bottom: a typical spheroidal galaxy with no disk component, the elliptcal galaxy M87, located at 60 million light years from us (credits in the picture).}} \\label{fig:feedback} \\end{figure} \\subsection{The theoretical framework} The formation of disk galaxies is one of the major unsolved problems of modern astrophysics. The basic theoretical framework states that disk galaxies arise from the gravitational collapse of a rotating protogalactic cloud of gas within the gravitational potential well of the dark halo$^6$. The gas cools via radiative processes during the collapse and eventually settles in centrifugal equilibrium at the center of the halo potential well forming a rotationally supported gas disk provided that some angular momentum is retained during the collapse $^7$. These ideas were developed two decades ago and they still constitute the backbone of disk galaxy formation models $^{8,9,10,11,12,13}$. What has changed dramatically since then is the cosmological context in which such idea is applied, which reflects the remarkable progress that cosmology has undergone in the meantime. After two decades of active debate there is now one cosmological paradigm according to which the energy density of the Universe is dominated by cold dark matter and a cosmological constant, while ordinary baryonic matter contributes only to a few percent level (an even smaller contribution is yielded by neutrinos)$^{14}$. This model is supported by observations of the large scale mass distribution in the Universe traced by galaxies themselves$^{15}$ and by the power spectrum of density fluctuations inferred from the cosmic microwave background radiation$^{14}$. Cold dark matter interacts only via gravity with itself and with ordinary baryonic matter, and is not subject to any dissipative force. The governing evolutionary equation for dark matter is the collisionless Boltzmann equation that describes a zero-pressure fluid, also termed a collisionless fluid$^2$. In this model, called $\\Lambda$CDM (CDM stands for \"cold dark matter\", while $\\Lambda$ is the cosmological constant required to explain the observed acceleration of the Universe) structure forms hierarchically in a bottom-up fashion, starting from the amplification via gravitational instability of primordial small density fluctuations in the dark matter$^{16,17}$. Because of the scale-free nature of gravity and the dissipationless nature of cold dark matter, in such a model one expects the formation of self-similar, ellipsoidal collapsed objects, dark matter halos, at all scales$^{18}$. The largest halos are the last to form$^{17}$. Also, direct three-dimensional simulations of structure formation in a CDM Universe predict that halos of any mass and size should contain a swarm of smaller halos, the so-called substructure$^{19,20}$, as shown in Figure 3. \\begin{figure} \\centering{ \\resizebox{14cm}{!}{\\includegraphics{S2T_vs_Mstar.ps}}\\\\ \\caption{The ratio between the mass of the stellar spheroid ($S$) and the sum of the mass of the stellar spheroid and the stellar disk ($T$) as a function of galaxy mass from a galaxy sample of the Sloan Digital Sky Surveys (SDSS)${^12}$ (2007 Blackwell Publishing Ltd). Red points with error bars show the median $S/T$ as a function of stellar mass together with the 10 and 90 percentiles of the distribution.}} \\label{fig:feedback} \\end{figure} The model predicts quantitatively the size and mass of the dark halo of a galaxy with a given measured rotational velocity (the rotational velocity of stars and gas probes the depth of the galaxy gravitational potential well). For a galaxy like our own Milky Way, for example, its observed rotational velocity of $\\sim 220$ km/s implies a halo mass of about $10^{12}$ solar masses and a halo radius of about $300$ kpc$^{21}$ (by comparison, the disk of our galaxy contains about $6 \\times 10^{10}$ solar masses of stars and about $10^{10}$ solar masses of gas$^2$). Numerical simulations of the growth of dark matter halos also predict that the radial density profiles of such halos diverge near the center and are well described by a power-law, $\\rho \\sim r^{-\\gamma}$ ($\\rho$ being the density and $r$ the spherically averaged radius of the halo), with $\\gamma=1-1.5$ $^{22, 23, 24, 25}$. Since dark matter dominates by mass over ordinary baryonic matter, gas collapses within such halos, eventually forming a galaxy, because it is pulled inward by their gravitational attraction rather than collapsing due to its own gravity. In this scenario there is no such a thing as a galaxy forming in isolation, rather structure, both dark and baryonic, builds up via continous accretion and merging of smaller systems containing a mixture of dark and baryonic matter$^{26}$. This highly dynamical picture emerging from the current cosmology is the main difference compared to earlier attempts to study galaxy formation. Halos gain their angular momentum by tidal torques due to asymmetries in the distribution of matter, and also by acquiring the angular momentum originally stored in their relative orbit as they come together and merge into a larger system$^{27,28,29}$. Prevailing models have then assumed that baryons and dark matter start with the same specific angular momentum before the collapse begins since they are subject to the same tidal torques $^{8}$. \\begin{figure} \\centering{ \\resizebox{16cm}{!}{\\includegraphics{fornaxlr.ps}}\\\\ \\caption{Example of a massive dark halo ($\\sim 10^{14}$ solar masses) assembled by hierarchical merging in a numerical simulation of the $\\Lambda$CDM structure formation model (simulation performed by L.Mayer and F.Governato with the GASOLINE code $^{44}$). Many small halos (\"substructure\") are orbiting inside the larger halo and will eventually merge with it as a result of dynamical friction eroding their orbital energy and orbital angular momentum. These dark matter lumps are the sites in which gas collapses and forms galaxies.}} \\label{fig:feedback} \\end{figure} \\subsection{Computer simulations:the angular momentum problem} The modern tool used to study the hierarchical growth of structure driven by gravity and the concurrent collapse of baryons within dark halos is represented by three-dimensional computer simulations that solve the gravitational and hydrodynamical forces between parcels of gas and dark matter. We will discuss the methodology employed by such simulations in the next section. For now it suffices to say that in the most popular simulation method both the gas (i.e. the baryonic component) and the dark matter are represented by particles so that structures are discretized in mass and space. The evolutionary equations, such as the collisionless Boltzmann equation for cold dark matter, the Euler equation for baryonic matter (baryonic matter is treated as an ideal gas) and the Poisson equation, which holds for both, are solved for such discrete representation of physical reality. Available methods to discretize physical variables and governing equations are constructed in such a way that they should converge to the exact continuum solution for an infinite number of particles. As we will see in the next section, discretization itself, along with other aspects of the current methods, can introduce spurious effects in the computer models. Simulations take advantage of large parallel supercomputers in which hundreds of processing units are used simultaneously to compute the forces and advance the system to the next timestep. One important prediction of simulations of a CDM Universe is that halos have a rather universal value of the angular momentum (per unit mass) at any given epoch quite irrespective of their precise mass assembly history. This is parameterized via the dimensionless spin parameter $\\lambda = J E^{1/2}/GM^{5/2}$, where $E$, $M$ and $J$ are the total energy, mass and angular momentum of the dark halo ($G$ is the gravitational constant). One can show that $\\lambda$ is proportional to the ratio between the rotational kinetic energy and the kinetic energy in disordered motions associated with the halo. Halos have a universal distribution of spin parameters, which peaks at a value $\\lambda \\sim 0.035$ $^{29}$. Simple spherical one-dimensional models that study disk formation in an isolated CDM halo (namely a halo that does not interact or merge with other halos) predict that the size of disks resulting from the infall and collapse of baryons matches the size of observed disks in galaxies very well$^8$. The models use mainly two inputs, both coming from cosmological simulations, the halo density profile, which is related to the gravitational pull that drives the gas collapse, and the initial specific angular momentum of gas as implied by the typical values of the spin parameter. They further assume that angular momentum is conserved during the collapse. Hence this result is simple and remarkable at the same time; it says that CDM halos have the right amount of angular momentum to form observed disk galaxies. For more than a decade researchers have tried to reproduce the latter result with fully three-dimensional computer simulations but have run into several problems. It was soon realized that, once the hypothesis of isolation is removed and hierarchical merging is accounted for, angular momentum can be lost by the gas to the dark matter due to a process known as dynamical friction$^{30,2,31}$. During mergers, previously collapsed clumps of gas and dark matter fall into a larger dark halo and suffer a drag force as they move through the latter. The loss of angular momentum caused by the drag force, called \"dynamical friction\", is more effective when the gas is distributed into cold and dense lumps rather than being smooth and extended$^{32}$. But gas is expected to be clumpy in a model with collisionless cold dark matter in which collapse can occur at all scales, and large halos grow by accreting smaller halos which bring their own dense collapsed gas. As a result, early simulations$^{32}$ were obtaining improbable small disks with ten times less angular momentum than real ones. Two types of solutions for this \"angular momentum problem\" have been considered since then. The first is very drastic and calls for revising the cosmological model itself. Alternative models in which the dark matter has a non-negligible thermal velocity rather then being \"cold\" would produce collapsed systems only above a characteristic scale because the thermal jittering will tend to smear out short-wavelength density perturbations$^{33}$. These warm dark matter models (WDM) behave like CDM on large scale, thus maintaining its succesful features. The reduced clumpiness of dark matter halos in the WDM model implies that baryons are smoothly distributed rather than arranged in previously collapsed dense lumps when they fall into large galaxy-sized halos, and therefore lose less angular momentum by dynamical friction$^{33,34}$. The second, less exotic possibility, is that baryons do not just follow the merging hierarchy imposed by dark matter but somehow decouple from it and remain much smoother. This could happen if the thermal energy content of baryons was enough to resist gravitational collapse, at least up to some critical mass scale. This way a fraction of the gas that would have entered a halo in dense clumps within smaller halos would instead enter with a smooth distribution, perhaps avoiding catastrophic loss of angular momentum. Various plausible astrophysical mechanisms can be responsible for increasing the thermal energy content of the baryons, for example the energy injection by supernovae explosions and the ambient radiation field produced by stars, accreting black holes or also external galaxies. There is, however, a third possibility. This is that the baryons clump excessively in computer simulations because the numerical methods adopted can introduce artificial loss of angular momentum. As we argue in this report, a solution lies probably in a combination of the latter two proposals, with no need of revising the standard cosmological structure formation model. \\begin{figure} \\centering{ \\resizebox{18cm}{!}{\\includegraphics{tobiaslr.ps}}\\\\ \\caption{ Disk size as a function of mass resolution in numerical simulations of disk formation in an isolated CDM halo$^{48}$ (2007 Blackwell Publishing Ltd). The three panels show density maps of gas in a slice through the centre of the gaseous galactic disk after 5 Gyr; the gas mass resolution decreases from the left to the right by about a factor 8 each snapshot (the maximum resolution is 1 million gas particles). The box side length is 20 kpc for every panel. At all resolutions disks show asymmetries such as central bar-like structures and spiral arms. The bar, however, is a strong and long-lived feature only for sufficiently high spatial resolution (set by the gravitational softening)$^{48}$.}} \\label{fig:feedback} \\end{figure} The next two sections will be devoted, respectively, to the role of numerical effects in disk formation simulations, and to the modeling of gas thermodynamics and star formation in the simulations. We will then show how the structure of simulated disks is affected by different models of thermodynamics and star formation. Finally, we will summarize the current status of the field and the major problems that remain to be solved, including the puzzling origin of disk galaxies without a bulge. We will attempt to recall the most important contributions by the various groups actively involved in this field of research while at the same time covering in more detail some recent results of the research group to which the authors of this report belong. ", "conclusions": "" }, "0801/0801.1071_arXiv.txt": { "abstract": "We summarize and discuss recent work (Fregeau 2007) that presents the confluence of three results suggesting that most Galactic globular clusters are still in the process of core contraction, and have not yet reached the thermal equilibrium phase driven by binary scattering interactions: that 1) the three clusters that appear to be overabundant in X-ray binaries per unit encounter frequency are observationally classified as ``core-collapsed,'' 2) recent numerical simulations of cluster evolution with primordial binaries show that structural parameters of clusters in the binary-burning phase agree only with ``core-collapsed'' clusters, and 3) a cluster in the binary-burning phase for the last few Gyr should have $\\sim 5$ times more dynamically formed X-ray sources than if it were in the core contraction phase for the same time. ", "introduction": "This proceedings article briefly summarizes my contributed talk at the ``Population Explosion'' meeting, which itself was a brief summary of the work described in \\citet{fregeau07}. I have taken the opportunity here to include supplemental material, but first refer the reader to the detailed, complete discussion in \\citet{fregeau07}. \\subsection{X-Ray Sources and Cluster Dynamics} Although X-ray binaries have been known to be overabundant in Galactic globular clusters (by unit mass relative to the disk) for over 30 years \\citep{1975ApJ...199L.143C,1975Natur.253..698K}, it is only recently that clear evidence for their dynamical origin has revealed itself. \\citet{2003ApJ...591L.131P} found that the number of X-ray sources with $L_X > 4 \\times 10^{30}\\,{\\rm erg}/{\\rm s}$, $N_X$, in a cluster follows a clear, nearly linear correlation with the encounter frequency $\\Gamma$, a measure of the dynamical interaction rate in the cluster. There are a few clear outliers to this trend, however. Terzan 1 is overabundant in $N_X$ by a factor of $\\sim 20$, NGC 6397 by a factor of $\\sim 5$, and NGC 7099 by a factor of $\\sim 2$ \\citep{2006MNRAS.369..407C,2007ApJ...657..286L}. The common thread among these clusters is that they are all classified observationally as ``core-collapsed,'' while the other clusters in the sample are not. A cluster is observationally termed core-collapsed if its surface brightness profile is consistent with a cusp at the limit of observational resolution. Since this observational classification is linked with the cluster evolutionary state (as described below), it appears that cluster evolution complicates the $N_X$--$\\Gamma$ correlation. \\subsection{Understanding Cluster Core Radii} The evolution of a globular cluster, being a bound self-gravitating system, is very similar to that of a star, and comprises three main phases. The cluster first ``core contracts'' on a relaxation timescale. Once the core density becomes large enough for binary scattering interactions to begin generating energy, the cluster settles into the ``binary-burning'' phase, in which the cluster's core properties remain roughly constant with time. Once the binary population is exhausted in the core, it collapses via the gravothermal instability, leading to extremely high central densities, followed by a series of gravothermal oscillations in which the core expands and contracts repeatedly. (For a graphical representation of the three main phases of cluster evolution, see Figure 1 of \\citet{1991ApJ...370..567G}, Figure 5 of \\citet{2003ApJ...593..772F}, or Figure 29.1 of \\citet{2003gmbp.book.....H}.) Recent comparison of star cluster evolution simulations with observations of cluster structural parameters have shown that the clusters observationally classified as core-collapsed are in fact most likely in the binary-burning phase, while the rest are most likely still in the core-contraction phase \\citep{2006MNRAS.368..677H,2007ApJ...658.1047F}. \\subsection{A Refined $\\Gamma$} First introduced by \\citet{1987IAUS..125..187V}, the encounter frequency $\\Gamma$ is an estimate of the {\\em current} dynamical interaction rate in the cluster, which is assumed to be proportional to the current number of observable X-ray sources. Typically, it is approximated as \\begin{equation}\\label{eq:standardgamma} \\Gamma \\equiv \\frac{dN_{\\rm int}}{dt} \\propto \\rho_c^2 r_c^3 / v_\\sigma \\, , \\end{equation} where $r_c$ is the cluster core radius, $\\rho_c$ is the core mass density, and $v_\\sigma$ is the core velocity dispersion. More accurate approximations of the interaction rate have been used, including numerical integrals over cluster models \\citep{2003ApJ...591L.131P}, but all are estimates of the {\\em current} interaction rate. Recent work has shown that dynamically formed X-ray binaries have finite detectable lifetimes ($\\sim 10^5$ to $\\sim 10^9$ yr, depending on binary type), and furthermore that there is a lag time of several Gyr between dynamical interaction and the binary turning on as an X-ray source \\citep{2006MNRAS.372.1043I,2007arXiv0706.4096I}. Since the binary interaction rate is a function of the cluster properties (density, velocity dispersion, etc.), it is clear that the X-ray binaries we see in clusters today were formed several Gyr ago, and thus encode the recent cluster evolution history in their populations. We thus adopt a refined version of $\\Gamma$ which encodes this history by integrating the interaction rate over the dynamical evolution of the cluster. The resulting quantity is the total number of strong dynamical interactions, which should be roughly proportional to the currently observable number of X-ray binaries. Plugging in the different phases of cluster evolution will yield different estimates for the number of sources, and possibly allow one to differentiate among the different phases by comparing with the observed number of sources. Using standard assumptions, and adopting some results from recent $N$-body simulations for the evolution of $r_c$ with time, one finds the ratio of the number of interactions for a cluster in the binary-burning phase to one in the core contraction phase is \\begin{equation} \\frac{N_{\\rm int,bb}}{N_{\\rm int,cc}} = \\frac{t_x}{t_0} \\frac{2.835}{\\left(\\frac{t_0}{t_0+9t_\\ell}\\right)^{0.315}- \\left(\\frac{t_0}{t_0+9t_\\ell+9t_x}\\right)^{0.315}} \\, , \\end{equation} where $t_x$ is the X-ray source lifetime, $t_\\ell$ is the lag time between strong dynamical interaction and X-ray turn-on, and $t_0$ is the current cluster age. (The gravothermal oscillation phase is excluded from the discussion since in this phase the core binary fraction would be much smaller than what is observed.) This expression has a minimum of $2.0$ and a maximum of $17.8$ in the range $t_x=10^{-4}$--$3\\,{\\rm Gyr}$, $t_\\ell=1$--$10\\,{\\rm Gyr}$, for $t_0=13\\,{\\rm Gyr}$. For the canonical values of $t_x=1\\,{\\rm Gyr}$ and $t_\\ell=3\\,{\\rm Gyr}$ with $t_0=13\\,{\\rm Gyr}$, the value is $5.0$. Since the number of X-ray sources should scale roughly linearly with the number of interactions, this suggests that if a cluster is in the binary-burning phase (and has been for a time $t_\\ell+t_x$ to the present), it should have $\\sim 5$ times as many X-ray sources than it would if it were in the core contraction phase. The three observationally core-collapsed clusters in the sample previously mentioned are overabundant in $N_X$ by factors compatible with the range of values allowed by this expression, suggesting that the observationally core-collapsed clusters are indeed in the binary-burning phase, while the rest are still in the process of core contraction. \\subsection{Discussion} Since only $\\sim 20$\\% of Galactic globular clusters are observationally classified as core-collapsed, the conclusion that seems strongly suggested is that most clusters ($\\sim 80$\\%) are currently still in the core contraction phase, while those that are core-collapsed are in the binary burning phase. This goes counter to the widely held belief that most clusters are currently in the binary burning phase, and complicates the many existing studies that have assumed cluster core properties that are constant with time, including predictions of blue straggler populations, tidal-capture binaries, and the evolution of the core binary fraction \\citep[e.g.,][]{2004ApJ...605L..29M,1994ApJ...423..274D,2005MNRAS.358..572I}. Similarly, this result does away with the need for alternate energy sources in cluster cores to explain currently observed core radii, including intermediate-mass black holes in most clusters, mass loss from stellar mergers, and ongoing mass segregation \\citep{2006astro.ph.12040T,chatterjeeposter,2004ApJ...608L..25M}. ", "conclusions": "A few interesting aspects of this analysis did not fit into the original {\\em letter}. It has been pointed out that an important but neglected factor going into the interaction rate is the neutron star retention fraction, since it depends on the cluster escape speed, and since it is neutron star $+$ stellar binary interactions that predominantly lead to X-ray binaries (Rasio, private communication). Since there is no reason to expect that the cluster escape speed at the time of neutron star formation does not vary from cluster to cluster, the correlation between dynamical interaction rate and $N_X$ should be blurred significantly by this effect. We note that the analysis in \\citet{fregeau07} avoids this issue entirely by comparing a cluster only with a version of itself in a different evolutionary phase. Another issue is that there is apparently a significant contribution to the X-ray binary population from primordial binaries in clusters with small encounter frequencies. (This is likely the reason for the sub-linear dependence of $N_X$ on $\\Gamma$ in \\citet{2003ApJ...591L.131P}.) This has given rise to the use of the interaction rate per unit cluster mass, $\\gamma \\equiv \\Gamma/M_{\\rm clus}$, to help isolate the dynamically-formed population in the analysis. The use of $\\gamma$ has, for example, shown that a significant fraction of cataclysmic variables are dynamically formed \\citep{2006ApJ...646L.143P}. The results of \\citet{fregeau07} could be further tested in the future with the more sensitive $\\gamma$ as the library of deep X-ray observations of globular clusters grows. Finally, we mention that in the near future we plan to test our simple overabundance prediction by using the detailed numerical models of \\citet{2005MNRAS.358..572I}. Although there are some hints from that paper that inputting a time-varying core does not noticeably affect the final core binary fraction, we expect that the number of visible X-ray binaries, while noisy, may be more noticeably affected." }, "0801/0801.4462_arXiv.txt": { "abstract": "Since GRBs fade rapidly, it is important to publish accurate, precise positions at early times. For \\Swift-detected bursts, the best promptly available position is most commonly the X-ray Telescope (XRT) position. We present two processes, developed by the \\Swift\\ team at Leicester, which are now routinely used to improve the precision and accuracy of the XRT positions reported by the \\Swift\\ team. Both methods, which are fully automated, make use of a PSF-fitting approach which accounts for the bad columns on the CCD. The first method yields positions with 90\\% error radii $<$4.4\" 90\\% of the time, within 10--20 minutes of the trigger. The second method astrometrically corrects the position using UVOT field stars and the known mapping between the XRT and UVOT detectors, yielding enhanced positions with 90\\% error radii of $<$2.8\" 90\\% of the time, usually ~2 hours after the trigger. ", "introduction": "For the majority of \\Swift-detected GRBs, the best promptly available position is that of the X-ray telescope (XRT, \\cite{Burrows05}). It is thus desirable to reduce the 3.5\" boresight uncertainty associated with this instrument. We have developed two techniques to achieve this goal. The first is a fitting technique which accounts for hot columns on the X-ray CCD. This is described in Section~1 and the application of this to promptly available data is detailed in Section~2. The second technique is applied to the full ground dataset, and uses the field stars in the UV/Optical telescope (UVOT, \\cite{Roming05}) to astrometrically correct the XRT position, eliminating the XRT's boresight uncertainty. This is described in Section~3. In Fig.~\\ref{fig:errdist} we show the distribution of position uncertainties produced by these techniques, comparing them with positions determined onboard the XRT, and the `refined' positions produced from the full dataset without astrometric correction. Finally, in Section~4 we discuss forthcoming improvements to the second technique, and the potential for applying it to the prompt data. For an overview of the different positions available from the \\Swift\\ XRT, see \\url{http://www.swift.ac.uk/xrt_pos.php} \\begin{figure} \\includegraphics[height=8cm,angle=-90]{errdist} \\includegraphics[height=8cm,angle=-90]{errcheck} \\caption{\\emph{Left:} Distribution of the 90\\% confidence error radii produced for XRT positions of GRBs. Distributions shows are those obtained onboard automatically (solid), from SPER data (Section 2, dotted), on the ground from the full dataset (short dashes) and the enhanced positions (Section 3, long dashes).\\emph{Right:} Distribution of the offsets of the UVOT-enhanced positions from the UVOT position for GRBs with both, divided by the position error. As can be seen, 90\\% of the enhanced positions agree with the UVOT positions, confirming the error circle is correctly calibrated.} \\label{fig:errdist} \\end{figure} ", "conclusions": "" }, "0801/0801.0594_arXiv.txt": { "abstract": "{} {We present $J$, $H$, $K$ spectrally dispersed interferometry with a spectral resolution of 35 for the Mira variable S~Orionis. We aim at measuring the diameter variation as a function of wavelength that is expected due to molecular layers lying above the continuum-forming photosphere. Our final goal is a better understanding of the pulsating atmosphere and its role in the mass-loss process.} {Visibility data of S~Ori were obtained at phase 0.78 with the VLTI/AMBER instrument using the fringe tracker FINITO at 29 spectral channels between 1.29\\,$\\mu$m and 2.32\\,$\\mu$m. Apparent uniform disk (UD) diameters were computed for each spectral channel. In addition, the visibility data were directly compared to predictions by recent self-excited dynamic model atmospheres.} {S~Ori shows significant variations in the visibility values as a function of spectral channel that can only be described by a clear variation in the apparent angular size with wavelength. The closure phase values are close to zero at all spectral channels, indicating the absence of asymmetric intensity features. The apparent UD angular diameter is smallest at about 1.3\\,$\\mu$m and 1.7\\,$\\mu$m and increases by a factor of $\\sim$\\,1.4 around 2.0\\,$\\mu$m. The minimum UD angular diameter at near-continuum wavelengths is $\\Theta_\\mathrm{UD}$=8.1$\\pm$0.5\\,mas, corresponding to $R\\sim$\\,420\\,R$_\\odot$. The S~Ori visibility data and the apparent UD variations can be explained reasonably well by a dynamic atmosphere model that includes molecular layers, particularly water vapor and CO. The best-fitting photospheric angular diameter of the model atmosphere is $\\Theta_\\mathrm{Phot}$=8.3$\\pm$0.2\\,mas, consistent with the UD diameter measured at near-continuum wavelengths.} {The measured visibility and UD diameter variations with wavelength resemble and generally confirm the predictions by recent dynamic model atmospheres. These size variations with wavelength can be understood as the effects from water vapor and CO layers lying above the continuum-forming photosphere. The major remaining differences between observations and model prediction are very likely due to an imperfect match of the phase and cycle combination between observation and available models.} ", "introduction": "Mira stars are low-mass, large-amplitude, long-period variable stars on the AGB, evolving toward the planetary nebula and white dwarf phases. They exhibit a mass-loss rate on the order of $\\sim$10$^{-6}$\\,M$_\\odot$/year that significantly affects the further stellar evolution and is one of the most important sources for the chemical enrichment of the interstellar medium. The dust condensation sequence, the wind-driving mechanism, and the role of pulsation are currently not well understood, in particular for oxygen-rich AGB stars (Woitke et al. \\cite{woitke06}; H\\\"ofner \\& Andersen \\cite{hoefner07}). The pulsating atmospheres of Mira stars can become very extended because of dynamic effects including shock fronts, and they are very cool in their outer parts. Here, molecules can form, which for O-rich stars are most importantly H$_2$O, CO, TiO, and SiO (Tsuji et al. \\cite{tsuji97}; Tej et al. \\cite{tej03}; Ohnaka \\cite{ohnaka04}). Wittkowski et al. (\\cite{wittkowski07}) found for the case of S~Ori that Al$_2$O$_3$ dust condenses within the extended atmosphere at phase-dependent distances of 1.8--2.4 photospheric radii. This extended atmosphere, which is characterized by phase-dependent temperature and density stratifications, the presence of molecular layers, and the formation of dust, is thus of particular interest for our better understanding of pulsation and mass loss. Observed radii of Mira stars have been found to differ for different optical and infrared bandpasses (e.g. Thompson et al. \\cite{thompson02}; Mennesson et al. \\cite{mennesson02}; Ireland et al. \\cite{ireland04a}; Perrin et al. \\cite{perrin04}; Eisner et al. \\cite{eisner07}), and this has been attributed to the presence of molecular layers located above the continuum-forming photosphere. Here, we present both a spectro-interferometric observation of the Mira star S~Ori that covers the near-infrared $J$, $H$, and $K$ bands simultaneously at a spectral resolution of 35 and a comparison to recent self-excited dynamic model atmospheres. S~Ori is a Mira variable star with spectral type M6.5e--M9.5e and $V$ magnitude 7.2--14.0 (Samus et al. \\cite{samus04}). We use a Julian Date of last maximum brightness $T_0=2453190\\ \\mathrm{days}$, a period $P=430\\ \\mathrm{days}$ (as in Wittkowski et al. \\cite{wittkowski07}), and the distance of $480\\ \\mathrm{pc}\\ \\pm\\ 120\\ \\mathrm{pc}$ from van Belle et al. (\\cite{vanbelle02}). The broadband near-infrared $K$ UD angular diameter of S~Ori has been measured by van Belle et al. (\\cite{vanbelle96}), Millan-Gabet et al. (\\cite{millan05}), and Boboltz \\& Wittkowski (\\cite{boboltz05}) to values between 9.6\\,mas and 10.5\\,mas at different phases. Joint VLTI/MIDI and VLBA/SiO maser observations by Wittkowski et al. (\\cite{wittkowski07}) have shown that S~Ori exhibits significant phase-dependencies of the atmospheric extension and dust shell parameters with photospheric angular diameters between 7.9\\,mas and 9.7\\,mas. \\begin{table*} \\caption{Observation log. Night starting 12 October 2007, JD 2454386.} \\label{tab:obslog} \\begin{tabular}{lllllllrrrr} \\hline\\hline Target & Purpose & $\\Theta_\\mathrm{LD}$& DIT & Time & $\\Phi_\\mathrm{Vis}$ & $B_p$ [m] & PA$_p$ & AM & Seeing & $\\tau_0$\\\\ & & [mas] &[msec] &[UTC] & & E0-G0/G0-H0/E0-H0 &$\\deg$ && [$\\arcsec$]&[msec]\\\\\\hline \\object{45~Eri} & Calibrator (K3 II-III) &2.15$\\pm$0.04& 25 & 08:03-08:07 & &16.0/31.9/47.9&-107&1.1&1.3&1.3\\\\ 45~Eri & Calibrator (K3 II-III) &2.15$\\pm$0.04& 50 & 08:09-08:12 & &16.0/32.0/48.0&-107&1.1&1.3&1.3\\\\ \\object{$\\gamma$~Eri}& Check star (M0.5 IIIb)&8.74$\\pm$0.09& 25 & 08:22-08:26 & &15.6/31.2/46.8&-104&1.1&1.4&1.2\\\\ $\\gamma$~Eri& Check star (M0.5 IIIb)&8.74$\\pm$0.09& 50 & 08:28-08:32 & &15.5/31.0/46.5&-104&1.1&1.4&1.2\\\\ S~Ori&Science target&& 25 & 08:45-08:49 & 0.78 &15.9/31.8/47.8&-107&1.1&1.3&1.3\\\\ S~Ori&Science target&& 50 & 08:52-08:57 & 0.78 &16.0/31.9/47.9&-107&1.1&1.3&1.3\\\\ $\\alpha$~Hor & Calibrator (K2 III) &2.76$\\pm$0.03& 25 & 09:14-09:18 &&14.5/28.9/43.3&-91&1.1&1.2&1.4\\\\ $\\alpha$~Hor & Calibrator (K2 III) &2.76$\\pm$0.03& 50 & 09:21-09-24 &&14.3/28.7/43.0&-90&1.1&1.4&1.2\\\\ S~Ori & Science target && 25 & 09:36-09:40 & 0.78 &15.9/31.8/47.7&-106&1.1&1.5&1.1\\\\\\hline \\end{tabular} \\end{table*} ", "conclusions": "" }, "0801/0801.2558_arXiv.txt": { "abstract": "I review measurements of star formation in nearby galaxies in the UV--to--FIR wavelength range, and discuss their impact on SFR determinations in intermediate and high redshift galaxy populations. Existing and upcoming facilities will enable precise cross-calibrations among the various indicators, thus bringing them onto a common scale. ", "introduction": "Determinations of star formation rates (SFRs) in galaxies utilize indicators at a variety of wavelengths, from the X--ray to the radio. Many indicators have been defined in response to specific needs. For instance, when new populations of galaxies are discovered using a new wavelength window or improved observing techniques/instruments in a certain waveband, there is a push to investigate whether that waveband can be used to derive SFRs as well, and/or to define the uncertainties and limitations of doing so. The advent of new facilities (e.g., Herschel, LMT, ALMA, JWST, etc., and the many ground--based telescopes under construction or design) together with existing ones (HST, Spitzer, Chandra, and the vast array of existing ground--based facilities) will cover extensively the electromagnetic spectrum at unprecedented sensitivities. This will offer the opportunity to cross-calibrate many of the SFR indicators across a range of redshifts, and, therefore, on many galaxy populations at various stages of their evolution. In this brief review, I discuss the current status and the known limitations of SFR indicators in a few wavelength regimes: ultraviolet (UV), optical/near--infrared, and mid/far--Infrared (MIR/FIR). ", "conclusions": "" }, "0801/0801.2591_arXiv.txt": { "abstract": "The discovery of over 200 extrasolar planets with the radial velocity (RV) technique has revealed that many giant planets have large eccentricities, in striking contrast with most of the planets in the solar system and prior theories of planet formation. The realization that many giant planets have large eccentricities raises a fundamental question: ``Do terrestrial-size planets of other stars typically have significantly eccentric orbits or nearly circular orbits like the Earth?'' Here, we demonstrate that photometric observations of transiting planets could be used to characterize the orbital eccentricities for individual transiting planets, as well the eccentricity distribution for various populations of transiting planets (e.g., those with a certain range of orbital periods or physical sizes). Such characterizations can provide valuable constraints on theories for the excitation of eccentricities and tidal dissipation. We outline the future prospects of the technique given the exciting prospects for future transit searches, such as those to be carried out by the CoRoT and Kepler missions. ", "introduction": "\\label{sec_intro} Theorists have proposed numerous mechanisms that could excite orbital eccentricities. Some of these mechanisms are expected to affect all planets independent of their mass (e.g., perturbations by binary companions, passing stars, or stellar jets; e.g., Holman et al.\\ 1997; Laughlin \\& Adams 1998; Ford et al.\\ 2000; Zakamska \\& Tremaine 2004; Namouni 2007), while the efficiency of other mechanisms would vary with planet mass (e.g., planet-disk or planet-planet interactions; e.g., Artymowicz 1992; Goldreich \\& Sari 2003; Chatterjee et al.\\ 2007). If the mechanism(s) exciting eccentricities of the known giant planets also affect terrestrial planets, then Earth-mass planets on nearly circular orbits could be quite rare. On the other hand, if large eccentricities are common only in systems with massive giant planets and/or very massive disks, then there may be an abundance of planetary systems with terrestrial planets on low eccentricity orbits (Beer et al.\\ 2004). Thus, understanding the eccentricity distribution of terrestrial planets could provide empirical constraints for planet formation theories (e.g., Ford \\& Rasio 2007) and shed light on the processes that determined the eccentricity evolution in our solar system. Since the discovery of transiting giant planets with eccentric orbits, authors have begun to consider the implications of eccentricities for transiting planets (e.g., Barnes 2007; Burke 2008). The CoRoT and Kepler space missions are expected to detect many transiting planets and measure their sizes and orbital periods, including some in or near the ``habitable zone'' (e.g., Kasting et al.\\ 1993). The Kepler mission aims to determine the frequency of Earth-like planets, and study how the frequency and properties of planets correlates with the properties of their host stars (Basri et al.\\ 2005). Since a significant eccentricity would cause the stellar flux incident on the planet's surface to vary, a planet's eccentricity affects its climate (i.e., equilibrium temperature, amplitude of seasonal variability) and potentially its habitability (Williams et al.\\ 2002; Gaidos \\& Williams 2004). Our method could be applied to these planets to determine the frequency of terrestrial planets that could also have Earth-like climates, and thus influence the design of future space missions that will attempt to detect and characterize nearly Earth-like planets (e.g., Space Interferometry Mission-PlanetQuest) and search them for signs of life (e.g., Terrestrial Planet Finder). In \\S\\ref{SecDuration}, we show how the duration of a transit is affected by a planet's orbital eccentricity. We outline how to interpret the transit duration for transiting planets with both low signal-to-noise light curves (\\S\\ref{SecLoSN}) and high signal-to-noise (\\S\\ref{SecHiSN}). We compare the magnitude of the effect on the transit duration to the expected precision of eccentricity constraints based on Kepler photometric data (\\S\\ref{SecStatErr}) and also the typical accuracy of stellar parameters (\\S\\ref{SecStarErr}). We demonstrate that Kepler observations could be used to characterize the orbital eccentricities of terrestrial planets. We describe statistical approaches for analyzing the distribution of transit durations of a population of transiting planets in \\S\\ref{SecStats}. In \\S\\ref{SecRV}, we discuss the role of radial velocity observations for constraining eccentricities of Earth-like transiting planets. In \\S\\ref{SecDiscuss}, we conclude with a discussion of how the results could contribute to understanding the formation and evolution of terrestrial planets and address fundamental questions, such as ``What is the frequency of terrestrial planets that have Earth-like eccentricities?'' and ``What is the frequency of terrestrial planets that pass through the habitable zone?'' ", "conclusions": "\\label{SecDiscuss} We have described how photometric observations can constrain the eccentricities of individual transiting planets and characterize the eccentricity distribution of a population of planets. For each planet, we can test the null hypothesis that it is on a circular orbit and calculate the posterior probability distribution for the eccentricity. A combination of such analyses for several transiting planets could be used to characterize the eccentricity distribution of a population of planets. For example, this method could be used to investigate how the fraction of eccentric orbits varies with the orbital period or physical proprieties of the star and planet. We expect this type of analysis to become increasingly powerful given the rapidly growing number of known transiting planets. This type of analysis will be particularly valuable for low-mass transiting planets or transiting planets around faint host stars. In both cases, radial velocity follow-up observations will be extremely challenging. This will be the case for many transiting planet candidates found by space-based transit searches, such as CoRoT and Kepler. The capability of these missions to discover terrestrial-mass planets is particularly exciting. Our method could could determine if terrestrial planets with low eccentricities like the Earth are common or rare. This would provide significant constraints on theories proposed to explain the eccentricities of extrasolar planets. For example, Kepler might find many terrestrial planets with low eccentricity orbits, suggesting that the mechanisms that excite the eccentricities of giant planets are often ineffective for terrestrial mass planets. In this scenario, those terrestrial planets that do have large eccentricities might typically be accompanied by nearby giant planets, suggesting that it is the giant planets are responsible for exciting the eccentricities of terrestrial planets (Veras \\& Armitage 2005, 2006). Alternatively, Kepler might find that terrestrial planets are much more common than giant planets, and yet they still commonly have large eccentricities. This could arise from eccentricity excitation mechanisms that do not require giant planets, or due to interactions with previous giant planets that have since been ejected, accreted or destroyed by the star (e.g., Ford et al.\\ 2005; Raymond et al.\\ 2006; Mandell et al.\\ 2007). Our method also provides a means for studying the tidal evolution of short-period planets. Tidal effects are likely to circularize planets with sufficiently short orbital periods. For short-period planets, we can compute the tidal circularization timescale ($t_{\\rm circ}$) based on the properties of the star and planet. If there is a sharp transition between circular and eccentric orbits, then this could be used to place constraints on tidal theory (e.g., Zahn \\& Bouchet 1989; Melo et al.\\ 2001; Mathieu et al.\\ 2004). For eccentric planets with relatively short $t_{\\rm circ}$, it may be possible to place a lower limit on the $Q$ factor the is related to the planet's internal structure (e.g., Ford et al.\\ 1999; Bodenheimber et al.\\ 2001; Maness et al.\\ 2007; Mardling 2007). In a Bayesian framework, one can calculate the posterior probability distribution for the fraction of each orbit during which the planet-star separation is between an inner and outer cut-off. If the cut-offs are set to be the putative boundary of the habitable zone, then we can then ask, ``What fraction of terrestrial planets are in the habitable zone for some/at least half/all of their orbit?''. The results of such investigations could have implications for the climates of potentially habitable planets, the frequency of such planets, and the design of future missions that aim to detect and characterize nearby Earth-like planets that could harbor life (e.g., Marcy et al.\\ 2005). We have demonstrated that it is practical to collect sufficient photons to characterize the eccentricity distribution of terrestrial extrasolar planets, but we assumed that limb-darkening parameters can be well constrained by some combination of stellar modeling and external observations. Our analytical estimates have not incorporated limb darkening effects or potential systematic uncertainties in stellar models. Future search should address both of these effects. Multi-wavelength observations (particularly in the infrared) could be particularly useful for addressing both these issues. In particular, we plan to investigate the potential for combinations of space-based detections and ground-based follow-up observations to improve the characterization of the eccentricities of transiting planets. Finally, we note that this method for characterizing the eccentricities of terrestrial planets from transit light curves underscores the importance of developing and validating precise and accurate stellar models. Uncertainties in stellar parameters models are expected to dominate the error budget for bright target stars. The potential for systematic uncertainties due to stellar modeling will make it particularly challenging to study low eccentricity systems as a function of stellar properties. Fortunately, we these concerns would not preclude our method from being applied to terrestrial-sized planets recognizing the relatively large eccentricities typical for giant planets." }, "0801/0801.0846_arXiv.txt": { "abstract": "We consider the bound states of the massive scalar field around a rotating black hole immersed in the asymptotically uniform magnetic field. In the regime of slow black hole rotation, the Klein-Gordon equation allows separation of variables. We show that the growth rate of the instability can be amplified a few times by the magnetic field. The effect occurs because the magnetic field adds the \"effective mass\" term $B |m|$ to the scalar field potential for a Kerr black hole. In addition, and as a by-product, we discuss the behavior of the quasinormal modes for the magnetized rotating black holes. ", "introduction": "Interaction of black holes and magnetic fields is an important factor for large astrophysical black holes, and, especially, for supermassive galactic black holes because of enormous magnetic fields in the nucleus of galaxies and near supermassive black holes. Strong magnetic fields are induced also by accretion disks of rotating charged matter near black holes. Yet, interaction of strong magnetic fields and black holes is, apparently, not limited just by the context of large scale astrophysical systems. Recent development in the brane world theories opens possibility of observing quantum gravity at the Tev scale, so that mini black holes might be created at particle collisions in the Large Hadron Collider. Observing black holes in laboratories could make it possible to test interaction of black holes with external fields. After all, in the early universe strong magnetic fields could effect primordial black holes. Sufficiently strong magnetic field near a black hole deforms the black hole space-time, so that one cannot consider the magnetic field as a test field on the black hole background, but a consistent solution of the Einstein-Maxwell equations is necessary instead. Fortunately, exact solutions for such a situation exist for both non-rotating (the Ernst solution) \\cite{8} and rotating cases (the Diaz solution) \\cite{9}. Yet, the essential question is whether such a strong magnetic field, that could deform the black hole space-time, can exist, and, whether the Ernst and Diaz solutions can be considered as at least some approximate approach to a realistic situation?. The Ernst solution is described by the metric \\begin{equation} d s^{2} = \\Lambda^{2} \\left( \\left(1- \\frac{2 M}{r} \\right) d t^{2} - \\left(1- \\frac{2 M}{r} \\right)^{-1} d r^{2} -r^{2} d \\theta^{2} \\right) - \\frac{r^{2} \\sin^{2} \\theta}{\\Lambda^{2} } d \\phi^{2}, \\end{equation} where the external magnetic field is determined by the parameter $B$, \\begin{equation} \\Lambda = 1 + \\frac{1}{4} B^{2} r^{2} \\sin^{2} \\theta, \\end{equation} and the \"unit\" magnetic field measured in $Gs$ is \\begin{equation} B_{M} = 1/M = 2.4 \\times 10^{19} \\frac{M_{Sun}}{M}. \\end{equation} The magnetic field in the Ernst solution is poloidal and homogeneous far from the black hole. This is not a big restriction for our consideration. Indeed, although the large scale magnetic field in the observable universe has both toroidal and poloidal components, it is the poloidal component that dominate in centers of galaxies and thereby in the region near a super-massive black hole. For mini black holes, which according to the brane-world scenarios, are expected to be observed in particles collisions, the homogeneous (poloidal) magnetic field is natural. Even though the super-strong accelerating magnetic field is screened in the region of particles collisions, we can assume, at least in principle, that if mini black holes are created in a laboratory, one could \"immerse\" mini-black holes in the magnetic field with required properties. In order to make estimations of possible influence of the magnetic field on the supermassive black holes we need the two parameters at hand: the magnetic field parameter $B$ and the mass of the black hole $M$. Modern observations suggest that super-massive black holes in centers of galaxies can have mass $M \\approx 10^6 - 10^9 M_{Sun}$, while observations of the magnetic field are much more complicated and lead sometimes to controversial predictions for the value of the magnetic field in the centers of galaxies \\cite{Zakharov:2002cf}, \\cite{Contra}. At the same time, observations of the Active Galactic Nuclei (AGN) and micro- quasars indicate the existence of wide X-ray emission of lines of heavy ionized elements in their spectra \\cite{Zakharov:2002cf}. This effect is very well described when supposing existence of the super-strong magnetic field $B \\approx 10^{10}- 10^{11}$ $Gs$ \\cite{Zakharov:2002cf}. In addition some other theories \\cite{Tolubaev} imply existence of a strong magnetic field in the AGN. From these data for $M$ and $B$ at hand, by formula (2), one can easily see that for the heaviest galactic black holes, the magnetic field can be of order of fractions of a \"unity\" ($1/M$), thereby effecting the black hole metric itself. The Diaz and Ernst solutions, which we shall analyze here, describe the black holes immersed in a magnetic field decaying near the black hole to some asymptotic value, so that far from black hole the magnetic field is uniform and represents the Melvin Universe. Different effects around such black holes have been studied in \\cite{10} and some generalizations were obtained in \\cite{11}. Potentially observable effects, such as quasinormal modes and gravitational lensing were considered in \\cite{12}, \\cite{13}, \\cite{14} for non-rotating magnetized black holes. In this paper, we shall consider scattering of scalar waves around rotating magnetized black holes. An incident wave near the black hole will be partially absorbed (tunneled through the potential barrier) and partially radiated away as a response of a black hole to an external perturbation. When a black hole of mass $M$ and radius $r_{+}$ is rotating, and the real oscillation frequency of the perturbation (for a mode with an azimuthal number $m$) satisfies the inequality \\begin{equation} \\omega < \\frac{a m}{2 M r_{+}}, \\end{equation} then the energy radiated away exceeds the energy of the incident wave. This is the well known effect of super-radiance predicted by Zeldovich and Misner \\cite{15} and computed for Kerr black holes by Starobinsky \\cite{16}. Press and Teukolsky suggested that if the wave radiated away will be reflected by a mirror surrounding a black hole, one could make a kind of \"black hole bomb\", because initial small perturbations would grow without bound \\cite{17}. The role of the mirror can play the potential well, another local minimum, which appears because of the non-vanishing massive term \\cite{18}. In the limit $M \\mu <<1$, where $M$ is the black hole mass and $\\mu$ is the inverse Compton wavelength of the particle, Detweiler showed that the massive scalar field exerts the superradiant instability with the maximal growth rate \\cite{20}, \\cite{21} (at $a=M$) given by the formula \\begin{equation} \\gamma = \\frac{1}{24} \\frac{a}{M} (\\mu M)^{9} (G M/c^{3})^{-1}. \\end{equation} This instability has very small growth rate and probably cannot be significant for real black holes. Accurate calculations of instability made recently in \\cite{18_1} allowed to observe the maximal growth \\begin{equation} \\gamma = 1.5 10^{-7} (G M/c^{3})^{-1}. \\end{equation} at $M \\mu =0.42$. Yet, due-to short lifetimes or large particle masses, this instability is tiny for any known massive boson particles, except possibly, the neutral pion $\\pi^{0}$ when $M \\sim 10^{12} kg.$ In this paper we shall consider the bound states of massive scalar field in the vicinity of the magnetized black holes and analyze the corresponding superradiant instability. As a by-product we shall consider the $B$-corrections to the orbital geodesic motion (as that for which the instability is important) of particles around the Ernst black hole. The paper is organized as follows. In Sec. II the massive scalar field equation is decoupled for the regime of \"small\" (in comparison with a \"unity\" $1/M$) magnetic field, and is reduced to the wave-like equation with an effective potential. Sec. III considers the bound states and their instability in the limit $\\mu M \\ll 1$. In particular, we obtain the generalization of the well-known Detweiler formula for the superradiant instability for a non-zero external magnetic field. In Sec. IV we briefly discuss the quasinormal modes of rotating magnetized black holes. In the Conclusion we summarize the obtained results and outline the number of open questions. \\vspace{4mm} ", "conclusions": "We have separated variables for the Klein-Gordon and Hamilton-Jacoby equations in the limit $B \\ll M^{-1}$. This corresponds to the \"weak\" magnetic field as for the deformation the black hole space-time, but a super-strong field in the astrophysical context. It is shown that the growth rate of the superradiant instability may be amplified by a few times due-to the magnetic field. This enlarges the range of black hole masses for which the instability may be significant in the neutral pion decay. The essential problem which was beyond our investigation is the calculation of the instability growth for the highly rotating black holes. Another important question is a detailed description of the process of filling of the quasi-bound states \\cite{Galtsov2}." }, "0801/0801.0900_arXiv.txt": { "abstract": "In this series we construct an effective field theory (EFT) in curved spacetime to study gravitational radiation and backreaction effects. We begin in this paper with a derivation of the self-force on a compact object moving in the background spacetime of a supermassive black hole. The EFT approach utilizes the disparity between two length scales, which in this problem are the size of the compact object $r_m$ and the radius of curvature of the background spacetime ${\\cal R}$ such that $\\varepsilon \\equiv r_m / {\\cal R} \\ll 1$, to treat the orbital dynamics of the compact object, described as an effective point particle, separately from its tidal deformations. The equation of motion of an effective relativistic point particle coupled to the gravitational waves generated by its motion in a curved background spacetime can be derived without making a slow motion or weak field approximation, as was assumed in earlier EFT treatment of post-Newtonian binaries. Ultraviolet divergences are regularized using Hadamard's {\\it partie finie} to isolate the non-local finite part from the quasi-local divergent part. The latter is constructed from a momentum space representation for the graviton retarded propagator and is evaluated using dimensional regularization in which only logarithmic divergences are relevant for renormalizing the parameters of the theory. As a first important application of this framework we explicitly derive the first order self-force given by Mino, Sasaki, Tanaka, Quinn and Wald. Going beyond the point particle approximation, to account for the finite size of the object, we demonstrate that for extreme mass ratio inspirals the motion of a compact object is affected by tidally induced moments at $O(\\varepsilon^4)$, in the form of an Effacement Principle. The relatively large radius-to-mass ratio of a white dwarf star allows for these effects to be enhanced until the white dwarf becomes tidally disrupted, a potentially $O(\\varepsilon^2)$ process, or plunges into the supermassive black hole. This work provides a new foundation for further exploration of higher order self force corrections, gravitational radiation and spinning compact objects. ", "introduction": "In two previous papers \\cite{GalleyHu:PRD72,GalleyHuLin:PRD74,Galley:PhD}, using a stochastic field theory approach based on open system concepts, we derive the scalar, electromagnetic and gravitational self-force to leading order on a particle moving in an arbitrary curved background spacetime. % We begin with the particle following a quantum mechanical path while interacting with a linear quantum field \\cite{JohnsonHu:PRD65,JohnsonHu:FoundPhys35,Johnson:PhD}. The conditions on a stochastic field theory (for open systems) to emerge from a quantum field theory (of closed systems) are that the mass and size of the particle are large enough so the particle worldline is sufficiently decohered from its interactions with the quantum fluctuations of the field that it can be considered as quasi-classical, and yet sufficiently small that quantum fluctuations manifest as classical stochastic forces \\cite{GellMannHartle:PRD47}. \\subsection{Quantum, Stochastic and Effective Field Theories} When there is a significant discrepancy between the two mass (or energy or length) scales in a problem, as in the extreme mass ratio binary systems under consideration here, one could use an open system stochastic description for their dynamics, such as developed in \\cite{GalleyHu:PRD72,GalleyHuLin:PRD74}. When this discrepancy is huge (such as between the QCD and GUT scales in particle physics) the stochastic component is strongly suppressed in the wide range between the two scales, away from the threshold region \\cite{CalzettaHu:PRD55}. Then the stochastic field theory description will give rise to an effective field theory (EFT) description \\cite{Burgess:EFT} for the motion of the small mass subsystem. Due to the large separation in the mass scales quantum loop corrections from the field and the intrinsic quantum mechanical worldline fluctuations are very strongly suppressed. These two factors render a quantum field theory (QFT) into a stochastic field theory (with sufficiently decohered histories) and in turn (with sufficiently small stochasticity) an effective field theory for the dynamics of the reduced systems. We shall explain the essence and demonstrate the advantages in taking a field theory approach to treat radiation-reaction of classical systems. The application of EFT to the treatment of gravitational radiation from post-Newtonian (PN) binary systems was first introduced by Goldberger and Rothstein \\cite{GoldbergerRothstein:PRD73}. Our formulation of a curved spacetime effective field theory (CS-EFT) goes beyond with two important features: it is for any curved spacetime background and there is no slow motion or weak field restrictions \\footnote{In addition to \\cite{GalleyHu:PRD72, GalleyHuLin:PRD74, GoldbergerRothstein:PRD73} for applying effective field theory techniques to the motion of extended bodies and charges, see also a recent paper of \\cite{KolSmolkin:grqc}.}. In this and three subsequent papers \\cite{Galley:EFT2, Galley:EFT3, Galley:spin} we construct a CS-EFT and apply it to derive the self-force on a compact object moving in an arbitrary curved background For concreteness, the background spacetime is assumed to be that of a supermassive black hole (SMBH) with curvature scale $\\cR$ much larger than the size of the compact object $r_m$. The smallness of the ratio $\\ve \\equiv r_m / \\cR$ makes it a good expansion parameter for a perturbation theory treatment describing the extreme mass ratio inspiral (EMRI) of the compact object. These binary systems are expected to be good candidates for detecting gravitational wave signatures using the space-based gravitational wave interferometer LISA \\cite{LISA}. We now give some realistic numbers for astrophysical processes in this category to delimit the range of validity for carrying out these perturbation expansions for these EMRI sources detectable in LISA's bandwidth . However, we emphasize that the formalism we construct here is of a sufficiently general nature that it can be applied to any compact object moving in an arbitrary curved background, including those spacetimes sourced by some form of stress-energy and those possessing a cosmological constant. \\subsection{Relevant scales in EMRIs} Consider the motion of a compact object (a black hole, neutron star or white dwarf with a mass $m$ ranging from about 1 to 100 solar masses) moving through the spacetime of a SMBH with a mass $M \\sim 10^{5-7} M_\\odot$. We have in mind that the compact object moves in a stationary background provided by the supermassive black hole, such as the Schwarzschild or Kerr spacetimes. Such spacetimes are appropriate for a description of the EMRI in which the compact object is bound by the gravitational pull of the SMBH. By emitting gravitational waves the binary system loses energy until the compact object plunges into the SMBH. The emission of gravitational radiation from such a system is expected to be detected with the anticipated construction and launch of the LISA space-based interferometer \\cite{LISA}. It is believed that most SMBHs lurking in the middle of galaxies, which are thought to host the prime sources of gravitational wave emissions detectable by LISA, are spinning and clean in the sense that most, if not all, of the surrounding material has already fallen into the black hole. (Active galactic nuclei are a notable exception \\cite{LISAsources}.) Because of this the Kerr background is perhaps the most astrophysically relevant spacetime for the extreme mass ratio inspiral. The Kerr solution is vacuous ($R_{\\mu\\nu}=0$), stationary and stable under small perturbations \\cite{Whiting:JMathPhys30} and possesses two Killing fields. The first Killing field ($\\xi^\\alpha$) is time-like and describes time-translation invariance everywhere outside of the ergoregion. The second ($\\psi^\\alpha$) is space-like and describes the axial symmetry of the spacetime. The Ernst \\cite{Ernst:JMathPhys17} and Preston-Poisson \\cite{PrestonPoisson:PRD74} geometries describe a black hole immersed in an external magnetic field. From an astrophysical viewpoint, the external magnetic fields that a black hole at the center of a galaxy experiences are relatively weak and unlikely to significantly affect the motion of the compact object until a very high order in the perturbation theory. There are two relevant length scales in EMRIs. The smaller scale is set by the size of the compact object itself, denoted $r_m$. For an astrophysical black hole its radius is $r_{bh} = 2 G_N m \\sim m/ m_{pl}^2$ where $m_{pl}^{-2} = 32\\pi G_N$ in units where $\\hbar = c= 1$ \\footnote{We follow the conventions of \\cite{MTW} so that the metric has mostly positive signature $(-,+,+,+)$.}. For a neutron star with a mass $\\approx 1.4 M_\\odot$ and a radius of $10-16$ km it follows that $r_{ns} \\approx 4.8 -7.7 \\, G_N m \\sim m/m_{pl}^2$. Therefore, it is to be expected that the size of the compact object, be it a black hole or a neutron star, is of the order of its mass \\footnote{See Section \\ref{sec:finsize} for more details concerning a white dwarf, which has a typical radius $r_{wd} \\sim 10^3 m / m_{pl}^2$.} . The second relevant scale is the radius of curvature of the background spacetime, $\\cR$. We take $\\cR$ to be the following curvature invariant \\begin{eqnarray} \\cR = \\big( R_{\\mu\\alpha \\nu \\beta} R^{\\mu\\alpha \\nu \\beta} \\big) ^{-1/4} . \\label{curvinvariant0} \\end{eqnarray} For a (possibly rotating) stationary SMBH the radius of curvature is \\begin{eqnarray} \\cR \\sim \\sqrt{ \\frac{ m_{pl}^2 r^3}{M} } \\end{eqnarray} where $r$ is the typical orbital distance for the compact object away from the central black hole. For example, $r$ is the geometric mean of the semi-major and semi-minor axes of a compact object in an inclined elliptical orbit. In an approximately circular orbit $r$ is the orbital radius and for a particle moving faster than the escape velocity $r$ is the impact parameter. In the strong field regime where $r \\sim M / m_{pl}^2$ the curvature scale is also $\\sim M / m_{pl}^2$ implying that $r/\\cR \\sim m / M$ whence a perturbative expansion in $\\ve$ is equivalent to one in $m/M$. The typical variation in time and space of the background is $\\gsim \\cR$. The wavelength $\\lambda$ of radiated metric perturbations from the compact object in a bound orbit is \\begin{eqnarray} \\lambda \\sim \\sqrt{ \\frac{ m_{pl}^2 r^3 }{ M } } \\sim \\cR , \\end{eqnarray} which shows that the wavelength of the gravitational waves does not provide a separate scale independently from $\\cR$. \\subsection{The CS-EFT approach: Issues and main features} The effective field theory approach was first introduced for the consideration of gravitational radiation from post-Newtonian binary systems in \\cite{GoldbergerRothstein:PRD73}, spinning compact objects in \\cite{Porto:PRD73, PortoRothstein:PRL97, Porto:grqc0701106, PortoRothstein:0712.2032}, and dissipative effects due to the absorption of gravitational waves in \\cite{GoldbergerRothstein:PRD73_2, Porto:0710.5150}. See \\cite{Goldberger:LesHouches, PortoSturani:grqc0701105, GoldbergerRothstein:GRG38} for introductory reviews. Let us call these theories PN-EFT: they are constructed to describe the motion of two slowly moving compact objects in a \\emph{flat background}. In particular, the compact objects are treated as effective \\emph{point particles}, the worldlines of which carry non-minimal operators describing the multipole moments from companion-induced tidal deformations as well as possible spin degrees of freedom and other intrinsic moments. Below we describe briefly some general features of EFT and the specific differences between our new CS-EFT approach and the existing PN-EFT. The use of point particles to source the metric perturbations about the flat background spacetime (note that the high frequency waves in a quantum description corresponds to massless spin-two particles, the gravitons, in flat space quantum field theory) prompts the appearance of divergences. Fortunately there exists a well-established bank of tools and techniques in quantum field theory for regularizing these divergences and renormalizing the parameters and coupling constants of the theory. A theory is renormalizable in the effective field theory sense if observables are calculated in the low energy limit: the divergences can always be absorbed into a renormalization of the coupling constants of the infinite number of non-minimal worldline operators. The use of dimensional regularization is particularly useful in effective field theories because the renormalization group equations are mass-independent for this scheme indicating that only logarithmic divergences contribute to the renormalization procedure \\footnote{The effective field theory approach has also been used in \\cite{Leibovich:NRGRtalk} to derive the radiation reaction force on an electrically charged extended body interacting with its own electromagnetic radiation, generalizing the Abraham, Lorenz and Dirac (ALD) equation for a point charge. It can also be derived from a stochastic field theory perspective, which contains EFT. See \\cite{JohnsonHu:PRD65, JohnsonHu:FoundPhys35, Johnson:PhD}.}. Our CS-EFT approach differs from this group of work in two ways. First, we work with an arbitrary \\emph{curved spacetime}. Second, we allow for the compact object to move with \\emph{relativistic speeds} in \\emph{strong field} regions of the background spacetime. The post-Newtonian effective field theory of \\cite{GoldbergerRothstein:PRD73} treats bodies moving slowly through a weak gravitational field. \\subsubsection{In-In formulation for real and causal equations of motion} Technically there are also fundamental differences. To derive real and causal equations of motion we emphasize the need to use the closed-time-path (CTP) integral formalism based on an in-in generating functional \\cite{Schwinger:JMathPhys2, Keldysh:JEPT20, ZhouSuHaoYu:PhysRep118, Jordan:PRD33, CalzettaHu:PRD35, CalzettaHu:PRD37} (`in' and `out' here refer to the initial and final vacua used to define the vacuum transition amplitude in the generating functional). The authors of \\cite{GoldbergerRothstein:PRD73} use the in-out formalism which is acceptable for field theories in a flat background spacetime since the in- and out-vacua are equivalent up to an irrelevant phase in that special case. In the presence of spacetime curvature, the difference becomes serious, as we will show in \\cite{Galley:EFT2} for the EMRI scenario. The in-out formalism can calculate matrix elements in scattering processes, but not expectation values of physical observables in real-time evolution. The equations of motion for the effective particle dynamics obtained from an in-out formulation are not generally causal. An initial value formulation of quantum field theory via the CTP generating functional is the only correct way for the description of the system's evolution -- it guarantees real and causal equations of motion for the particle dynamics \\cite{Jordan:PRD33, CalzettaHu:PRD37}. \\subsubsection{Effective point particle description} Another important ingredient in our construction is an effective point particle description for the motion of the compact object. Going beyond the point particle approximation is necessary to include the effects of tidal deformations induced by the background curvature as well as the effects from spin and other intrinsic moments. Following \\cite{GoldbergerRothstein:PRD73}, we introduce all possible terms into the point particle action that are consistent with general coordinate invariance and reparameterization invariance (and invariance under $SO(3)$ rotations for a non-spinning spherically symmetric compact object). By implementing a matching procedure using coordinate invariant observables we can match the observables of the effective point particle theory with the long wavelength limit of observables in the full ``microscopic\" theory to determine the values of the coupling constants of the non-minimal terms. As we will show in Section \\ref{sec:finsize} this allows us to deduce the order at which finite size effects affect the motion of the compact object through the statement of an \\emph{Effacement Principle}. To our knowledge this has not been explicitly given in the literature before for the EMRI scenario. \\subsubsection{The power counting rules} Power counting is a generalization of dimensional analysis. In our perturbative treatment it is crucial for determining how the Feynman rules scale with the parameter $\\ve$. Once the scaling of the Feynman rules are known we determine all of the tree-level Feynman diagrams that appear at a particular order. Those diagrams containing graviton loops are safely ignored. We also assemble the diagrams that include the non-minimal worldline operators describing the finite size of the compact object. Significantly, this allows us to determine the order in $\\ve$ at which finite size effects enter the particle equations of motion. With the power counting rules the EFT approach becomes an efficient and systematic framework for calculating the self-force to any order in perturbation theory. Furthermore, by knowing how each Feynman diagram scales with $\\ve$ we can study a particular physical interaction that is of interest by focusing our attention on a single diagram or on a few diagrams without having to calculate every contribution that appears at that order and at lower orders. For example, the leading order spin-spin interaction (spin here refers to that of the compact object, not of the SMBH) contributes to the self-force at third order in $\\ve$ for a maximally rotating compact object and can be calculated from the appropriate Feynman diagram \\cite{Galley:PhD, Galley:spin}. The power counting rules, the Feynman rules and their scaling with $\\ve$ are derived in Sections \\ref{ch3:powercounting} and \\ref{ch3:Feynmanrules}. \\subsubsection{Divergences and Regularization} In Section \\ref{ch3:reg} we propose a method for regularizing the divergences that appear in the effective action. Our approach utilizes a mixture of distributional and momentum space techniques within the context of dimensional regularization. We know from previous work % that the finite part of the self-force is generally non-local and history dependent. However, the ultraviolet divergences are quasi-local and independent of the history of the effective point particle's motion. To isolate the quasi-local divergence from the non-local finite part we use the method of Hadamard's {\\it partie finie}, or finite part, from distribution theory. (See Appendix \\ref{app:distro} for a brief review of the definitions and concepts of distribution theory relevant in this work.) After isolating the divergence from the non-local, finite remainder we then use a momentum space representation for the propagator in a curved background, first derived for a scalar field by Bunch and Parker \\cite{BunchParker:PRD20}, to calculate the divergent contributions. Their method is straightforward but not efficient for higher spin fields, including gravitons in a curved space. (See also the related work of \\cite{HuOConnor:PRD30}.) A novel method applicable for any tensor field is developed in \\cite{Galley:momentum} for computing the momentum space representation of the Feynman propagator. The method is sufficiently general to do the same for any quantum two-point function, including the retarded propagator $D^{ret}_{\\alpha \\beta \\gamma^\\prime \\delta^\\prime} (x,x^\\prime)$. This approach makes use of diagrammatic techniques borrowed from perturbative quantum field theory. In Riemann normal coordinates, we expand the field action in terms of the displacement from the point $x$. The series can be represented in terms of Feynman diagrams, which allows for an efficient evaluation of each term in the expansion. Furthermore, we prove that some of the diagrams are zero to all orders. This identity is not recognized in \\cite{BunchParker:PRD20} even though its relation to certain steps made in their calculations is evident. \\subsection{MST-QW Equation} Assembling all these essential ingredients, in Section \\ref{sec:EFTMST-QW} we give a demonstration of how the curved spacetime effective field theory construction is implemented, outline the steps in the regularization of divergences, perform an actual calculation of the effective action and from it derive the equation of motion for the compact object including the effect of gravitational self-force. As an application we work to first order in the mass ratio (i.e., $\\ve$) and obtain the well-known Mino-Sasaki-Tanaka-Quinn-Wald (MST-QW) equation \\cite{MinoSasakiTanaka:PRD55, QuinnWald:PRD56}. This also sets the stage for calculating the second order self-force, the emitted gravitational waves and the motion of compact objects with spin \\cite{Galley:EFT2, Galley:EFT3, Galley:spin}. ", "conclusions": "We develop an effective field theory approach for systematically deriving the self-force on a compact object moving in an arbitrary curved spacetime without the slow motion or weak field restrictions. The EFT is a realization of the open quantum system paradigm in systems with a large scale separation such that the system's induced fluctuations from the backreaction of the coarse-grained quantum field is utterly negligible \\cite{CalzettaHu:PRD55}. An initial value formulation of quantum field theory is adopted here using the closed-time-path (CTP) formalism for the in-in generating functional, which guarantees real and causal equations of motion for the compact object. As an illustration of the procedures involved in our approach we showed how to derive the MST-QW equation describing the (first order) self-force on a compact object. The CTP formalism is needed for a calculation of the second order self-force as will be shown in \\cite{Galley:EFT2}. We describe the compact object as an effective point particle that is capable of accounting for tidally induced finite size effects. % In calculating the effective action we encounter ultraviolet divergences stemming from a point particle interacting with arbitrarily high frequency modes of a graviton field. Using Hadamard's {\\it partie finie} to isolate the non-local finite part from the quasi-local divergences we are able to implement dimensional regularization within a (quasi-local) momentum space representation for the graviton propagator \\cite{Galley:momentum}. As such, all power divergences can be immediately set to zero implying that only logarithmic divergences are relevant for renormalizing the parameters of the theory. At first order, the effective action has a power divergence and may therefore be trivially regularized using dimensional regularization. In the spirit of an Effacement Principle we find that the finite size of the compact object first affects its motion at $O(\\ve^4)$ for a non-spinning black hole and neutron star. For a white dwarf star we deduce that such effects may be enhanced until the white dwarf is tidally disrupted at $O(\\ve^2)$ in which case the effective point particle description, and in particular the effective field theory developed here, breaks down. One may conceivably construct a new effective field theory by treating the supermassive black hole, the white dwarf and the accreting mass as an effective point particle possessing many relevant non-minimal couplings to the background geometry describing the intrinsic moments of this composite object. The leading order finite size corrections cause a deviation from the motion of a minimally coupled point particle that is not caused by interactions with gravitons but is due to the torques that develop on the tidally deformed compact object. On the other hand, the self-force is affected by the induced moments of the compact object at $O(\\ve^5)$. In summary, the EFT approach has at least two major advantages over the existing approaches: It provides a systematic procedure for carrying out a perturbative treatment, and an economical way to treat the ultraviolet divergences. Our CS-EFT improves on the PN-EFT introduced in \\cite{GoldbergerRothstein:PRD73} in that it is valid for a general curved spacetime and not limited to slow motion or weak field conditions. These will prove to be of special benefit for higher order self-force calculations. We will apply these steps to calculate the self-force at second order in $\\ve$ \\cite{Galley:EFT2}, the gravitational radiation emitted by EMRIs \\cite{Galley:EFT3} and the self-force on spinning compact objects \\cite{Galley:spin}." }, "0801/0801.0307_arXiv.txt": { "abstract": "{\\bf Abstract.} There is accumulating evidence that (fundamental) scalar fields may exist in Nature. The gravitational collapse of such a boson cloud would lead to a {\\em boson star} (BS) as a new type of a compact object. Similarly as for white dwarfs and neutron stars, there exists a limiting mass, below which a BS is {\\em stable} against complete gravitational collapse to a black hole. According to the form of the self-interaction of the basic constituents and the spacetime symmetry, we can distinguish mini-, axidilaton, soliton, charged, oscillating and rotating BSs. Their compactness prevents a Newtonian approximation, however, modifications of general relativity, as in the case of Jordan-Brans-Dicke theory as a low energy limit of strings, would provide them with {\\em gravitational memory}. In general, a BS is a compact, {\\em completely regular} configuration with structured layers due to the anisotropy of scalar matter, an exponentially decreasing 'halo', a critical mass inversely proportional to constituent mass, an effective radius, and a large particle number. Due to the Heisenberg principle, there exists a completely stable branch, and as a coherent state, it allows for rotating solutions with {\\em quantised angular momentum}. In this review, we concentrate on the fascinating possibilities of detecting the various subtypes of (excited) BSs: Possible signals include gravitational redshift and (micro-)lensing, emission of gravitational waves, or, in the case of a giant BS, its dark matter contribution to the rotation curves of galactic halos. ", "introduction": "In this review, we will assume that {\\em fundamental scalar fields} exist in Nature, that --- in the early stages of the universe --- they would have formed absolutely stable soliton-type configurations kept together by their self-generated gravitational field. Theoretically, such configurations are known as {\\em boson stars} (BSs). In some characteristics, they resemble neutron stars; in other aspects, they are different and, thereby, astronomers may have a chance to distinguish BS signals from other compact objects, like neutron stars or black holes (BHs). BSs can be regarded as descendants of self-gravitating photonic configurations called {\\em geons} (gravitational electromagnetic units) and proposed in 1955 by Wheeler \\cite{Wh55}. In this review, we shall mainly concentrate on the results which are not included in the first reviews from 1992 \\cite{Je92,LP92,LM92} as well as later in the Marcel Grossman meetings in Jerusalem and Rome \\cite{mielkeMG8,MS02}. Since then, the number of publications investigating how BSs could possibly be detected have increased; we intend to focus strongly on these more recent research results. This could provide astronomers a handling on possible signals of BSs. All of these theoretical results are only a humble beginning of the understanding where and in which scenarios a BS could be detected. We hope other researchers may find this review stimulating for gaining new ideas or for refining older research. A second intention of this review is to distinguish clearly the different theoretical models underlying the label BS which could lead to different observational consequences. The first distinction is that the matter part of a BS can be described by either a complex or by a real scalar field. Then, there can be different interactions: (a) self-interactions described by scalar field potentials, (b) minimal coupling to gauge fields, the scalar field can be carrier of a charge (electric or hypercharge, e.g.). Moreover, the BS scalar field can couple, in standard general relativity (GR), minimally, or, in scalar-tensor (ST) or Jordan-Brans-Dicke (JBD) theory, non-minimally to gravity. In the latter case, if the strength of gravitational force is influenced by the JBD scalar field, there is an interaction of the BS scalar field with the real JBD scalar field; this can lead to {\\em gravitational memory} effects in BSs. JBD theory is closely related to low energy limits of superstring theories \\cite{CJ78} which imply the primordial production of scalar fields. Then, relics like the dilaton, the axion, combined as {\\em axidilaton}, or other moduli fields could remain in our present epoch as candidates. Already the first two papers on BSs provided the two main directions: The history of these {\\em hypothetical} stars starts in 1968 with the work of Kaup \\cite{K68} using a complex massive scalar field with gravitational interactions in a semi-classical manner. The energy-momentum tensor is calculated classically providing the source for gravitation; a very detailed investigation of the solution classes was done in \\cite{FLP87a}. However, one year later, Ruffini and Bonazzola \\cite{RB69} used field quantisation of a real scalar field and considered the ground state of $N$ particles. The vacuum expectation value of the field operators yield the same energy-momentum tensor and thus, not surprisingly, the same field equations. The two different physical constituents, complex versus quantised real scalar field BS, yield the same macroscopical results. It should be noted, however, that the gravitational field $g_{\\mu\\nu}$ is kept {\\em classical} due to the non-renormalisability of standard GR, or, alternatively, treated it as principal low energy part of some renormalisable superstring model. More recently, a BS using a field quantised complex scalar field has been constructed. BSs consisting of pure semi-classical real scalars cannot exist because their static solutions in flat spacetime are unstable due to Derrick's theorem \\cite{De64}, solutions may arise only if the real scalar field possesses a time-dependence leading to non-static {\\em oscillating} BSs. However, if a fermion star is present as well, a real scalar component can be added and if it interacts with the fermions, then a combined boson-fermion star is the result as in the first calculation by T.D.~Lee and Pang \\cite{LP87}; cf.~Section \\ref{leepang}. A challenge for the BS model is that, so far, no {\\em fundamental} scalar particle has been detected with certainty in the laboratory. Several are proposed by theory: The Higgs particle $h$ is a necessary ingredient of the standard model. However, the possible discovery of the Higgs boson of mass $m_{\\rm h}=114.5$ GeV/$c^2$ at the Large Electron Positron (LEP) collider at CERN \\cite{Ac00,Ba00} gives the BS strand of investigation a fresh impetus. For the BS, we need a scalar particle which does not decay; or if it decays (like the Higgs, e.g.), one has to assume that, in the gravitational binding, the inverse process is efficient enough for an equilibrium, as is the case for the $\\beta$-decay inside a neutron star. In the latter case, the direct physical predecessor of that kind of a BS is not clear; but, of course, unstable Higgs particles could not have formed a BS by themselves. In order to explore that unknown particle regime, one can play with the model parameters such as particle mass or interaction constants. Then, BSs of rather different sizes can occur: it could be just a `gravitational atom'; it could be as massive as the presumed BH in the central part of a galaxy; or it could be an alternative explanation for parts of the dark matter in the halo of galaxies. In 1995, experiments \\cite{AK02} proved the existence of the fifth possible state of matter, the {\\em Bose-Einstein condensate} (BEC); in 2001, the Nobel prize was awarded for its experimental realization in traps. BSs, if they exist, would be an astrophysical realization with a self-generated gravitational confinement, cf.~\\cite{JB01,BLV01}. Let us sum up some of the properties of complex scalar field BSs (following an earlier version of \\cite{T02}) in Table I as we shall discuss in Section III. \\begin{center} \\parbox{14cm} {TABLE I. Overview of some complex scalar field BS properties.} \\end{center} $$\\vbox{\\offinterlineskip \\hrule \\halign{&\\vrule#&\\strut\\quad\\hfil#\\quad\\hfil\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & {\\bf Property} && {\\bf BS} &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr \\noalign{\\hrule} height2pt&\\omit&&\\omit&&\\omit&\\cr & Constituents \\hfill && Scalars (Bose-Einstein-Condensation) \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Pressure support \\hfill && Heisenberg's uncertainty relation \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Size \\hfill && Gigantic up to very compact (few Schwarzschild radii): Table IV, Fig.~\\ref{fig2} \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Surface \\hfill && Atmosphere \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Appearance \\hfill && Transparent (if only gravitationally interacting) \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Structure \\hfill && Einstein-Klein-Gordon equation \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Gravitational potential \\hfill && Newtonian weak up to highly relativistic \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Last stable orbit \\hfill && None \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Rotation \\hfill && Differentially; discrete \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Avoiding a baryonic BH by \\hfill && Jet, particle dynamics \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Avoiding a scalar BH by \\hfill && Stability, evolution \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Gravitational redshift \\hfill && Comparable to neutron star values, but larger due to transparency \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Gravitational (micro-)lensing \\hfill && Extreme deflection angles possible (Fig.~\\ref{fig3}); MACHOS \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Luminosity \\hfill && Larger than luminosity of a BH \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Star disruption (tidal radius) \\hfill && Yes \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr & Distinctive observational \\hfill && Broadening of emission lines \\hfill &\\cr & \\ \\ \\ signatures \\hfill && Gravitational waves \\hfill &\\cr & \\hfill && {\\v C}erenkov radiation \\hfill &\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr height2pt&\\omit&&\\omit&&\\omit&\\cr} \\hrule}$$ \\subsection{Fundamental scalar fields in Nature?} The physical nature of the spin-0-particles out of which the BS is presumed to consist, is still an open issue. Until now, no fundamental elementary scalar particle has been detected with certainty in accelerator experiments, which could serve as the main constituent of the BS. In the theory of Glashow, Weinberg, and Salam, a Higgs boson doublet $(\\Phi^+, \\Phi^0)$ and its anti-doublet $(\\Phi^-, \\bar \\Phi^0)$ are necessary ingredients in order to generate masses for the `heavy photons', i.e.~the $W^{\\pm }$ and $Z^0$ gauge vector bosons \\cite{Qu83}. After symmetry breaking, only one real scalar particle, the Higgs particle $h:=(\\Phi^0 + \\bar \\Phi^0)/\\sqrt{2}$, remains free and occurs in the state of a constant scalar field background \\cite{Pe97}. As it is indicated by the rather heavy top quark \\cite{Ab95} of 176 GeV/$c^2$, the mass of the Higgs particle is expected to be below 1000 GeV/$c^2$. This is supported by the possible discovery of the Higgs boson of mass $m_{\\rm h}=114.5$ GeV/$c^2$ at the Large Electron Positron (LEP) collider at CERN \\cite{Ac00,Ba00}. In order to stabilise such a light Higgs against quantum fluctuations, a {\\em supersymmetric} extension of the standard model is desirable. However, then it will be accompanied by additional heavy Higgs fields $H^0$, $A^0$, and a charged doublet $H^{\\pm}$ in the mass range of 100 GeV/$c^2$ to 1 TeV/$c^2$. For the unlike case of a Higgs mass $m_{\\rm h}$ above 1.2 TeV/$c^2$, the self-interaction $U(\\Phi)$ of the Higgs field is so large that any perturbative approach of the standard model becomes unreliable. Tentatively, a conformal extension of the standard model with gravity included has been analysed, cf.~\\cite{PR94,HMMN95}. Future high-energy experiments at the LHC at {\\sc Cern} should reveal if Higgs particles really exist in Nature. As free particles, the Higgs boson of Glashow-Weinberg-Salam theory is unstable, e.g.~with respect to the decays $h \\rightarrow W^+ + W^-$ and $h \\rightarrow Z^0+ Z^0$, if it is heavier than the gauge bosons. In a compact object like the BS, these decay channels are expected to be in partial equilibrium with the inverse processes $Z^0+ Z^0 \\rightarrow h$ or, via virtual $W$ triangle graph, $Z^0 \\rightarrow h + \\gamma$, for instance \\cite{CL84}, by utilising gravitational binding energy. This is presumably in full analogy with the neutron star \\cite{HWW94,ST83,We99} or quark star \\cite{KWWG95,GKW95}, where one finds an equilibrium of $\\beta $- and inverse $\\beta $-decay or of a quark-gluon plasma (with similar decay and inverse-decay processes) and thus stability of the macroscopic star with respect to radioactive decay. The known spin-0-particles in particle physics, although effectively described by a Klein-Gordon (KG) equation, are not elementary, they consist of quarks having spin 1/2. Nevertheless, they possess some physical properties which can help us to understand which kind of particles can occur in the different BS types. The electrically charged pions $\\pi^{\\pm}$ with a mass of about 140 MeV/$c^2$ are a typical example for complex scalar fields \\cite{IZ80,MS93}. The electrically neutral ${\\rm K}^0$, $\\bar{\\rm K}^0$ mesons are described by a complex KG field as well; ${\\rm K}^0$ and $\\bar{\\rm K}^0$ are distinguished by their hypercharge $Y:=B+S =\\pm 1$, where $B=0$ is the baryon number and $S$ is the strangeness corresponding to a global $U(1)$ symmetry. In both cases, particles (e.g.~$\\pi^{+}$, ${\\rm K}^0$) and anti-particles (e.g.~$\\pi^{-}$, $\\bar{\\rm K}^0$) occur. Generally, Noether's theorem leads to a conserved charge \\be Q = e ( N^+ - N^- ) \\label{Ncharge} \\ee related to the number $N$ of particles and anti-particles, respectively, where $e$ defines the absolute value of the charge. Let us stress that if the KG equation for a complex scalar field admits a {\\em global} $U(1)$ symmetry, as in the case for the ${\\rm K}^0$, $\\bar{\\rm K}^0$ mesons, ``merely'' a conserved particle number $N$ arises and the charge is the total strangeness $S$. Regarding the high degree of instability of pions, $K$ mesons, and Higgs fields, the effect of gravitational binding is essential to see if these particles had enough time to form a BS. In string theories, there may exist several fundamental scalar particles having different global or local charges. For example, for complex scalar particles with global $U(1)$ symmetry or for real scalar particles, the BS charge shall be called just {\\em boson number} $N$. For complex scalar particles with local $U(1)$ symmetry, the BS has a charge given by (\\ref{Ncharge}). In a BS with temperature near zero, we will assume that all constituents are particles, i.e.~no anti-particles present, or vice versa. In his model, Kaup \\cite{K68} used a complex scalar field with global $U(1)$ symmetry so that no gauge bosons are present. Ruffini and Bonazzola \\cite{RB69}, however, took a real scalar field for the general relativistic BS and a complex scalar field with global $U(1)$ symmetry for the Newtonian BS description. In the case that $U(1)$ symmetry is {\\em local}, the complex scalar field couples to the gauge field contribution, cf.~\\cite{IZ80}, p.~31; Jetzer and van der Bij \\cite{JB89} called their hypothetical astrophysical object {\\em charged boson star} (CBS). Within the Glashow-Weinberg-Salam theory before symmetry-breaking of the group $SU(2)\\times U(1)$, this $U(1)$ charge describes the weak hypercharge, and not an electric one. During the earliest stages of the universe, a complex scalar field with different kinds of $U(1)$ charge (than electric or hyperweak) could have been generated. There is also the particle/field classification with respect to the spatial reflection $P$. Both a real and a complex scalar field can be either a {\\em scalar} or a {\\em pseudo-scalar}. Thus, particle characterisation will lead to different consequences for the detection of each BS model. \\subsection{Boson star as a self-gravitating Bose-Einstein condensate \\label{BEC}} Since Einstein and Bose it is well-known that scalar fields represent identical particles which can occupy the same ground state. Such a {\\em Bose-Einstein condensate} (BEC) has been experimentally realized in 1995 for cold atoms of even number of electrons, protons, and neutrons, see Anglin and Ketterle \\cite{AK02} for a recent review. In the mean-field ansatz, the interaction of the atoms in a dilute gas is approximated by the effective potential \\be U(|\\Psi|)_{\\rm eff}= \\frac{\\lambda}{4} |\\Psi|^4 \\, . \\label{Ueff} \\ee This leads to a {\\em nonlinear} Schr\\\"odinger equation for $\\Psi$, in this context known as Gross-Pitaevskii equation. In a microscopic approach, one introduces bosonic creation and annihilation operators $b^\\dagger$ and $b$, respectively, satisfying \\be [b,b^\\dagger] =1 \\ee and finds that every number conserving normal ordered correlation function $\\langle b_1 \\cdots b_n \\rangle$ splits into the sum of all possible products of contractions $\\langle b_i b_j \\rangle$ as in Wick's theorem of quantum field theory (QFT). For $n=1$ one recovers the Gross-Pitaevskii equation, whereas the next order leads to the Hartree-Fock-Bogoliubov equations, see \\cite{KB01} for details. Therefore, it is gratifying to note that BSs with {\\em repulsive} self-interaction $U(|\\Phi|)$ considered already in Refs.~\\cite{MS81,CSW86} have their counterparts in the effective potential (\\ref{Ueff}) of BEC. Thus, some authors \\cite{JB01,BLV01} advocate consideration of a cold BS as a {\\em self-gravitating} BEC on an astrophysical scale. Recently, the so-called {\\em vortices}, collective excitations of BECs with angular momentum in the direction of the vortex axis $z$, have been predicted \\cite{WH00} and then experimentally prepared \\cite{MCW00}. Since, quantum-mechanically, the z-component $J_z= a N \\hbar$ of the total angular momentum is necessarily quantised by the {\\em azimuthal} quantum number $a$, it is not possible to ``continuously\" deform this state to the ground state, and by this circumstance, contributing to its (meta-) stability. In 1996, axisymmetric solutions of the Einstein-KG equations have been found by us \\cite{SM96,MS96,SM98,SM99} which are rotating BSs (differentially due to the frame-dragging in curved spacetime) and exhibit, as a collective state, the same relation $J =aN$ for the total angular momentum as in the case of the vortices of an BEC and a meta-stability against the decay into the ground state as well; cf.~Section \\ref{rotaBS}. ", "conclusions": "" }, "0801/0801.4304_arXiv.txt": { "abstract": "We present an analysis of the 5--8~$\\mu$m \\textit{Spitzer}-IRS spectra of a sample of 68 local Ultraluminous Infrared Galaxies (ULIRGs). Our diagnostic technique allows a clear separation of the active galactic nucleus (AGN) and starburst (SB) components in the observed mid-IR emission, and a simple analytic model provides a \\textit{quantitative} estimate of the AGN/starburst contribution to the bolometric luminosity. We show that AGNs are $\\sim30$ times brighter at 6~$\\mu$m than starbursts with the same bolometric luminosity, so that even faint AGNs can be detected. Star formation events are confirmed as the dominant power source for extreme infrared activity, since $\\sim85\\%$ of ULIRG luminosity arises from the SB component. Nonetheless an AGN is present in the majority (46/68) of our sources. ", "introduction": "Ultraluminous Infrared Galaxies (ULIRGs, $L_\\mathit{IR}>10^{12} L_\\odot$) are the local counterparts of the high-redshift objects dominating the cosmic background in the far-infrared and millimetric bands. Unveiling the nature of their energy source is fundamental in order to understand the star formation history and the obscured AGN activity in the distant Universe. Since their discovery, several infrared indicators have been proposed to determine whether the central engine in ULIRGs is an AGN or a starburst (SB). The presence of high-ionization lines in the mid-IR spectra of ULIRGs points to AGN activity, while intense PAH emission features are typical of starburst environments (Genzel et al. 1998; Laurent et al. 2000). Recently, the absorption feature of amorphous silicate grains centered at 9.7~$\\mu$m has also been used together with the PAH emission to assess the nature of the obscured power source (Spoon et al. 2007). An alternate way to disentangle AGNs and SBs in ULIRGs has been proposed by Risaliti et al. (2006, hereafter R06), based on the separation of the two continuum components in 3--4~$\\mu$m spectra. This method has been successfully applied to a sample of $\\sim$50 nearby ULIRGs (Risaliti, Imanishi \\& Sani 2007, \\textit{submitted}) and provided an estimate of the average AGN/SB contribution to ULIRGs. The key reason for using the continuum emission at $\\lambda \\simeq$3--4~$\\mu$m as a diagnostic is the difference between the 3-$\\mu$m to bolometric ratios in AGNs and starbursts ($\\sim$two orders of magnitude larger in the former). This makes the detection of the AGN component possible even when the AGN is heavily obscured and/or bolometrically weak compared to the starburst. However the original prescription is limited by the low quality of the available L-band spectra of ULIRGs (R06, Imanishi et al. 2006), which makes the results on individual sources highly uncertain, except for the $\\sim$10--15 brightest objects. At present we have extended the analysis to the 5--8~$\\mu$m spectral band, using the observations of the IRS instrument (Houck et al.~2004) onboard \\textit{Spitzer}. We disentangled the AGN and SB contributions to the observed 5--8~$\\mu$m emission of ULIRGs by combining average spectral templates representing the different properties of the two physical processes at work. The high quality of \\textit{Spitzer}-IRS data, in addition to the relatively low dispersion of the intrinsic continuum properties of both AGNs and starbursts in this spectral range, allows a much more accurate determination of the AGN/SB components than possible at other wavelengths (e.g. X-rays) or with other diagnostic methods based on emission lines. In this paper we present our decomposition method, and discuss a simple analytical model providing a \\textit{quantitative} estimate of the AGN/SB contribution to the bolometric luminosity of each source. ", "conclusions": "The use of average templates for AGN and SB emission has allowed us to disentangle the two components in the 5--8~$\\mu$m spectra of 68 local ULIRGs, observed with the \\textit{Spitzer Space Telescope}. We have been able to detect an AGN in more than 60\\% of our sources, and estimate its contribution to the bolometric luminosity. In a statistical sense, we confirm that local Ultraluminous Infrared Galaxies are powered for $\\sim$85\\% by intense star formation and for the remaining $\\sim$15\\% by AGN activity. Our method proves to be successful in unveiling an intrinsically faint or obscured AGN inside a ULIRG. In this context we also put on a sound basis our initial assumption that the wavelength interval 5--8~$\\mu$m is an appropriate spectral range in order to search for AGNs: an AGN turns out to be approximately 30 times more luminous at 6~$\\mu$m than a starburst with the same bolometric luminosity." }, "0801/0801.4903_arXiv.txt": { "abstract": "This paper proposes a new phenomenology for strong incompressible MHD turbulence with nonzero cross helicity. This phenomenology is then developed into a quantitative Fokker-Planck model that describes the time evolution of the anisotropic power spectra of the fluctuations propagating parallel and anti-parallel to the background magnetic field~$\\bm{B}_0$. It is found that in steady state the power spectra of the magnetic field and total energy are steeper than $k_\\perp^{-5/3}$ and become increasingly steep as $C/{\\cal E}$ increases, where $C=\\displaystyle \\int d^3\\! x \\; \\bm{v}\\cdot\\bm{B}$ is the cross helicity, ${\\cal E}$ is the fluctuation energy, and $k_\\perp$ is the wavevector component perpendicular to~$\\bm{B}_0$. Increasing $C$ with fixed ${\\cal E}$ increases the time required for energy to cascade to smaller scales, reduces the cascade power, and increases the anisotropy of the small-scale fluctuations. The implications of these results for the solar wind and solar corona are discussed in some detail. ", "introduction": "Much of our current understanding of incompressible magnetohydrodynamic (MHD) turbulence has its roots in the pioneering work of Iroshnikov~(1963) and Kraichnan~(1965). These studies emphasized the important fact that Alfv\\'en waves travelling in the same direction along a background magnetic field do not interact with one another and explained how one can think of the cascade of energy to small scales as resulting from collisions between oppositely directed Alfv\\'en wave packets. They also argued that in the absence of a mean magnetic field, the magnetic field of the energy-containing eddies at scale~$L$ affects fluctuations on scales~$\\ll L$ much in the same way as would a truly uniform mean magnetic field. Another foundation of our current understanding is the finding that MHD turbulence is inherently anisotropic. Montgomery \\& Turner~(1981) and Shebalin, Matthaeus, \\& Montgomery~(1983) showed that a strong uniform mean magnetic field~$\\bm{B}_0$ inhibits the cascade of energy to small scales measured in the direction parallel to~$\\bm{B}_0$. This early work was substantially elaborated upon by Higdon~(1984), Goldreich \\& Sridhar~(1995, 1997), Montgomery \\& Matthaeus~(1995), Ng \\& Bhattacharjee~(1996, 1997), Galtier et al~(2000), Cho \\& Lazarian~(2003), Oughton et al~(2006), and many others. For example, Cho \\& Vishniac (2000) used numerical simulations to show that when the fluctuating magnetic field~$\\delta B$ is $\\ga B_0$ the small-scale turbulent eddies become elongated along the local magnetic field direction. Goldreich \\& Sridhar (1995) introduced the important and influential idea of ``critical balance,'' which holds that at each scale the linear wave period for the bulk of the fluctuation energy is comparable to the time for the fluctuation energy to cascade to smaller scales. Goldreich \\& Sridhar (1995, 1997), Maron \\& Goldreich~(2001), and Lithwick \\& Goldreich~(2001) clarified a number of important physical processes in anisotropic MHD turbulence and used the concept of critical balance to determine the ratio of the dimensions of turbulent eddies in the directions parallel and perpendicular to the local magnetic field. Over the last several years, research on MHD turbulence has been proceeding along several different lines. For example, one group of studies has attempted to determine the power spectrum, intermittency, and anisotropy of strong incompressible MHD turbulence using direct numerical simulations. (See, e.g., Cho \\& Vishniac 2000, M\\\"uller \\& Biskamp 2000, Maron \\& Goldreich 2001, Cho et~al~2002, Haugen et~al~2004, Muller \\& Grappin~2005, Mininni \\& Pouquet 2007, Perez \\& Boldyrev 2008). Another series of papers has explored the properties of anisotropic turbulence in weakly collisional magnetized plasmas using gyrokinetics, a low-frequency expansion of the Vlasov equation that averages over the gyromotion of the particles. (Howes et al~2006, 2007a, 2007b; Schekochihin et al 2007). These authors investigated the transition between the Alfv\\'en-wave cascade and a kinetic-Alfv\\'en-wave cascade at length scales of order the proton gyroradius~$\\rho_i$, as well as the physics of energy dissipation and entropy production in the low-collisionality regime. Turbulence at scales~$\\lesssim \\rho_i$ has also been examined both numerically and analytically within the framework of fluid models, in particular Hall MHD and electron MHD. (Biskamp, Schwarz, \\& Drake 1996, Biskamp et~al~1999, Matthaeus et al~2003; Galtier \\& Bhattacharjee 2003, 2005; Cho \\& Lazarian 2004; Brodin et al~2006, Shukla et al~2006). Another group of studies has investigated the power spectrum, intermittency, and decay time of compressible MHD turbulence. (Oughton et al 1995, Stone et~al~1998, Lithwick \\& Goldreich~2001, Boldyrev et al~2002, Padoan et al~2004, Elmegreen \\& Scalo 2004). Additional work by Kuznetsov (2001), Cho \\& Lazarian (2002, 2003), Chandran (2005), and Luo \\& Melrose (2006) has begun to address the way in which Alfv\\'en waves, fast magnetosonic waves, and slow magnetosonic waves interact in compressible weak MHD turbulence. Another recent development is the finding that strong incompressible MHD turbulence leads to alternating patches of alignment and anti-alignment between the velocity and magnetic-field fluctuations. (Boldyrev 2005, 2006; Beresnyak \\& Lazarian 2006; Mason, Cattaneo, \\& Boldyrev 2006; Matthaeus et~al~2007) These studies examined how the degree of local alignment (and anti-alignment) depends upon length scale, as well as the effects of alignment upon the energy cascade time and the power spectrum of the turbulence. The topic addressed in this paper is the role of cross helicity in incompressible MHD turbulence. The cross helicity is defined as \\begin{equation} C = \\int d^3\\!x \\; \\bm{v}\\cdot {B}, \\label{eq:defC} \\end{equation} where $\\bm{v}$ is the velocity and~$\\bm{B}$ is the magnetic field. The cross helicity is conserved in the absence of dissipation and can be thought of as a measure of the linkages between lines of vorticity and magnetic field lines, both of which are frozen to the fluid flow in the absence of dissipation (Moffatt 1978). In the presence of a background magnetic field, $\\bm{B}_0 = B_0 \\hat{z}$, the cross helicity is also a measure of the difference between the energy of fluctuations travelling in the $-z$ and $+ z$ directions. Dobrowolny, Mangeney, \\& Veltri~(1980) showed that MHD turbulence with cross helicity decays to a maximally aligned state, with $\\delta \\bm{v} = \\pm \\delta\\bm{ B}/\\sqrt{4\\pi\\rho}$, where $\\delta \\bm{v}$ and $\\delta \\bm{B}$ are the fluctuating velocity and magnetic field and $\\rho$ is the mass density. Different decay rates for the energy and cross helicity were also demonstrated by Matthaeus \\& Montgomery~(1980). In another early study, Grappin, Pouquet, \\& L\\'eorat~(1983) used a statistical closure, the eddy-damped quasi-normal Markovian (EDQNM) approximation, to study strong 3D incompressible MHD turbulence with cross helicity, assuming isotropic power spectra. They found that when~$C\\neq 0$, the total energy spectrum is steeper than the isotropic Iroshnikov-Kraichnan $k^{-3/2}$~spectrum. Pouquet et~al~(1988) found a similar result in direct numerical simulations of 2D incompressible MHD turbulence. More recently, Lithwick, Goldreich, \\& Sridhar~(2007) and Beresnyak \\& Lazarian (2007) addressed the role of cross helicity in strong MHD turbulence taking into account the effects of anisotropy. This paper presents a new phenomenology for strong, anisotropic, incompressible MHD turbulence with nonzero cross helicity, and is organized as follows. Section~\\ref{sec:wpc} presents some relevant theoretical background. Section~\\ref{sec:theory} introduces the new phenomenology as well as two nonlinear advection-diffusion equations that model the time evolution of the power spectra. Analytic and numerical solutions to this equation in the weak-turbulence and strong-turbulence regimes are presented in Sections~\\ref{sec:weak} and~\\ref{sec:strong}. Section~\\ref{sec:strong} also presents a simple phenomenological derivation of the power spectra and anisotropy in strong MHD turbulence. Section~\\ref{sec:trans} presents a numerical solution to the advection-diffusion equation that shows the smooth transition between the weak and strong turbulence regimes. Section~\\ref{sec:unequal} addresses the case in which the parallel correlation lengths of waves propagating in opposite directions along the background magnetic field are unequal at the outer scale. In Section~\\ref{sec:solarwind}, the proposed phenomenology is applied to turbulence in the solar wind and solar corona, and in Section~\\ref{sec:comp} the results of this work are compared to the recent studies of Lithwick, Goldreich, \\& Sridhar~(2007) and Beresnyak \\& Lazarian~(2007). ", "conclusions": "This paper proposes a new phenomenology for strong, anisotropic, incompressible MHD turbulence with cross helicity and introduces a nonlinear advection-diffusion equation [equation~(\\ref{eq:FPpm})] to describe the time evolution of the anisotropic power spectra of the $w^+$ and $w^-$ fluctuations. It is found that in steady state the one-dimensional power spectra of the energetically dominant $w^+$ fluctuations, $E^+(k_\\perp)$, is steeper than $k_\\perp^{-5/3}$, and that $E^+(k_\\perp)$ becomes increasingly steep as the fractional cross helicity~$\\sigma_c$ increases. Increasing $\\sigma_c$ also increases the energy cascade time of the $w^+$ fluctuations, reduces the turbulent heating power for a fixed fluctuation energy, and increases the anisotropy of the fluctuations at small scales. Although most of the discussion has focused on forced, steady-state turbulence, the results of this paper can also be applied to decaying turbulence. For example, equations~(\\ref{eq:tau1}) and (\\ref{eq:tau2}) can be used to estimate the time scale for turbulence to decay. The resulting prediction is that if the fluctuations are initially excited with $w_{k_f}^+ \\gg w_{k_f}^-$ and with comparable parallel correlation lengths at the outer scale, then the turbulence will decay into a state in which~$w^-$ fluctuations are absent, as in the earlier work of Dobrowolny, Mangeney, \\& Veltri (1980), Grappin et~al~(1983), and Lithwick \\& Goldreich (2003). This ``maximally aligned'' state will then be free from nonlinear interactions, and will persist for long times until it damps via linear dissipation." }, "0801/0801.4418_arXiv.txt": { "abstract": "We present an analysis of the Suzaku observation of the northeastern rim of the Cygnus Loop supernova remnant. The high detection efficiency together with the high spectral resolution of the Suzaku X-ray CCD camera enables us to detect highly-ionized C and N emission lines from the Cygnus Loop. Given the significant plasma structure within the Suzaku field of view, we selected the softest region based on ROSAT observations. The Suzaku spectral data are well characterized by a two-component non-equilibrium ionization model with different best-fit values for both the electron temperature and ionization timescale. Abundances of C to Fe are all depleted to typically 0.23 times solar with the exception of O. The abundance of O is relatively depleted by an additional factor of two compared with other heavy elements. We found that the resonance-line-scattering optical depth for the intense resonance lines of O is significant and, whereas the optical depth for other resonance lines is not as significant, it still needs to be taken into account for accurate abundance determination. ", "introduction": "X-ray spectroscopic imaging observations of supernova remnants (SNRs) enable us to investigate the thermodynamic state of the hot plasma they contain. The spatial structure of electron temperature ($kT_{\\rm e}$), ionization timescale, and elemental abundance have now been derived for many SNRs with ASCA, Chandra, XMM-Newton, and Suzaku. The X-ray emission from SNRs is usually assumed to be optically thin since the electron density is typically low (0.1$-$10\\,cm$^{-3}$). This is safely the case for continuum emission, however the optical depth for the resonance lines cannot be assumed to be negligibly small for bright SNRs, as first pointed out by Kaastra \\& Mewe (1995) in the context of X-ray observations. Resonance transitions are optically allowed transitions from the ground state of an ion. In the coronal limit, they are more likely to occur than other transitions, as there is a large population of ions in the ground state. If there is a sufficient ion column density along a particular line of sight, resonance line photons can be scattered out of that line of sight to appear at another location. Resonance line scattering, thus, leads to an underestimate of elemental abundances as well as biases in $kT_{\\rm e}$ determined by line intensity ratios. It should be noted that any random or turbulent velocities will tend to decrease the effects of resonance line scattering. The Cygnus Loop is one of the brightest SNRs in the soft X-ray band and appears as a rim-brightened shell. The Cygnus Loop is nearby (540\\,pc; Blair et al.~2005) and has a low neutral H column density (several $\\times 10^{20}$\\,cm$^{-2}$) and large apparent size (2.$\\!\\!^\\circ$5$\\times$3.$\\!\\!^\\circ$5; Levenson et al.~1997), which enable us to study its spatially-resolved soft X-ray emission. Analysis of IUE observations of the Cygnus Loop indicates a high shock velocity (130\\,km s$^{-1}$), departures from steady flow behind the shock, and significant optical depths for the UV resonance lines (Raymond et al.~1981). Subsequent studies (e.g., Raymond et al.~1988; Cornett et al.~1992) have shown that optical depth effects play an important role in shaping the UV and optical spectral and morphological properties of the Cygnus Loop. In this paper, we report on the very soft X-ray emission of the northeastern region of the Cygnus Loop as observed with the Suzaku observatory (Mitsuda et al.~2007). The extended low energy range of the Suzaku X-ray CCD camera (XIS; Koyama et al.~2007) combined with their superior energy resolution allows us to detect highly ionized C and N emission lines from the northeastern region of the Cygnus Loop (Miyata et al.~2007; paper I, hereafter). These authors determined the abundances of C, N, and O independently for the first time for this source and found that their relative abundances are roughly consistent with solar values. Throughout the present paper, CGS units are used unless specified otherwise. ", "conclusions": "We analyzed the very soft X-ray emission from the northeastern region of the Cygnus Loop with the Suzaku Observatory. The X-ray emitting plasma requires at least two spectral components; a high $kT_{\\rm e}$ component with small ionization timescale and a low $kT_{\\rm e}$ component with large ionization timescale. Usually, X-ray emitting plasmas of SNRs are assumed to be optically thin because the typical density is quite low. Since the region we analyzed is very bright with a potentially large line-of-sight path length, however, this assumption may not be valid for some resonance lines. The line-center cross section of resonance scattering can be expressed as \\begin{equation}% \\sigma = {\\sqrt{\\pi} e^2 \\over m c}{f \\over \\nu}\\left({v \\over c}\\right)^{-1} \\simeq 1.86 \\times 10^{-9}\\,{f \\over E}\\,v^{-1} \\quad\\mbox{cm}^2\\ , \\end{equation} where $f$ and $\\nu$ are the oscillator strength and frequency, respectively, of the line transition concerned, $v$ is the root-mean-square kinetic velocity of the ion, $m$ is the electron mass and other quantities have their usual meanings. In the last expression, $E$ is the line energy in keV. The kinetic velocity is composed of thermal and turbulent motions, as \\begin{equation}% v^2 = \\left({2kT_{\\rm i} \\over m_{\\rm i}}\\right)^2 + \\xi^2\\ , \\end{equation} where $kT_{\\rm i}$ is the ion temperature in keV, $m_{\\rm i}$ is the ion mass, and $\\xi$ represents the root-mean-square turbulent velocity due to motions other than thermal one. In the following analysis, we assume that the second term (turbulent motion) is negligibly small compared to the first one (thermal motion) in equation (2); the validity is discussed later. The line-center optical depth for resonance scattering is given by \\begin{equation}% \\tau = n_z \\sigma L = \\left({n_z \\over n_Z}\\right) \\left({n_Z \\over n_{\\rm H}}\\right) \\left({n_{\\rm H} \\over n_{\\rm e}}\\right)\\,n_{\\rm e} \\sigma L \\ , \\end{equation} where $L$ is the path length through the plasma, $n_z$ the ion density, $n_Z$ the element density, $n_{\\rm H}$ the hydrogen density, and therefore $n_z/n_Z$ represents the ionic fraction, $n_Z/n_{\\rm H}$ the elemental abundance relative to hydrogen. We assume $L = 2.5$\\,pc and the emission volume to be 2.5$\\times$0.93$\\times$2.5\\,pc$^3$. The electron density is obtained from the emission measure to be 1.25 and 1.35\\,cm$^{-3}$ for the high and low $kT_{\\rm e}$ components, respectively, and $n_{\\rm e}/n_{\\rm H}$ is taken to be 1.2. For the ion temperature, we assume $kT_{\\rm i} = kT_{\\rm e}$ = 0.236\\,keV, which is the density-weighted mean value from table~\\ref{table:vnei}. In the calculation of $\\sigma$ for the He-like K$\\alpha$ lines, we summed the oscillator strengths for the three main lines of the He-like triplet, namely the forbidden ($z$), intercombination ($x+y$), and resonance ($w$) lines. The ionic fractions of O$\\;${\\scriptsize\\rmfamily{VII}}\\,K$\\alpha$ and O$\\;${\\scriptsize\\rmfamily{VIII}}\\,K$\\alpha$ were calculated by taking into account the NEI condition for each component separately (using Masai 1984). The density-weighted overall ionic fractions of O$\\;${\\scriptsize\\rmfamily{VII}}\\,K$\\alpha$ and O$\\;${\\scriptsize\\rmfamily{VIII}}\\,K$\\alpha$ were then determined to be 0.51 and 0.33, respectively. The inferred values of $\\tau$ for O$\\;${\\scriptsize\\rmfamily{VII}}\\,K$\\alpha$ and O$\\;${\\scriptsize\\rmfamily{VIII}}\\,K$\\alpha$ turn out to be 0.54 and 0.16. Table~\\ref{table:optical_depth} summarizes the resonance-line-scattering cross sections and optical depths for the K emission lines detected in the Suzaku data and shown in table~\\ref{table:gaus_fit}. The optical depth of O$\\;${\\scriptsize\\rmfamily{VII}}\\,K$\\alpha$ is the largest whereas those of the other emission lines are not negligibly small. In order to model the effect of self-absorption in a simple way, we employed the so-called ``escape-factor'' method (Irons 1979). We assume the global (source- and direction-average) Doppler-profile escape factor for plane-parallel ``slab'' geometry (Bhatia \\& Kastner 2000) which is a reasonable assumption for the shell-brightened appearance of the Cygnus Loop. Escape factor values for the relevant emission lines are summarized in table~\\ref{table:optical_depth}. All values are less than 0.87, indicating that our line intensities have been underestimated by a factor of 13\\% or more. Optical depth effects, therefore, play a significant role in the X-ray emission from the Cygnus Loop with the O abundance, in particular, being underestimated by a factor of 20--40\\%. However, optical depth effects alone cannot account for the entire abundance discrepancy of a factor of two that we observe. It should be noted as well that the escape factor is different for O$\\;${\\scriptsize\\rmfamily{VII}}\\,K$\\alpha$ and O$\\;${\\scriptsize\\rmfamily{VIII}}\\,K$\\alpha$, implying that the {\\it true} line intensity ratio differs from the observed value. Correcting the line intensity ratio of O$\\;${\\scriptsize\\rmfamily{VIII}}\\,K$\\alpha$ to O$\\;${\\scriptsize\\rmfamily{VII}}\\,K$\\alpha$ for this effect results in a revised ratio of 0.21, which implies $kT_{\\rm e}=0.15$\\,keV based on the APEC model of equilibrium ionization. This optical depth effect not only reduces derived elemental abundances but it also biases fitted $kT_{\\rm e}$ measurements from their {\\it true} values. Furthermore, varying the elemental abundance affects the calculation of the optical depth as shown in equation (3). Future work will be needed to develop more sophisticated emission codes that take optical depth effects into account. We have made a few assumptions that tend to enhance the effects of resonance line scattering, including (a) temperature equipartition between electrons and ions, (b) no bulk motions, and (c) no turbulence. The temperature equipartition is a reasonable assumption for the region observed, as argued by Ghavamian et al. (2001) based on the optical spectra as well as by Miyata \\& Tsunemi (1999) on X-rays. Bulk motions may not be such a big problem for the Cygnus Loop due to its relatively low shock speed (few hundred km\\,s$^{-1}$) and its large angular size on the sky. The region observed is right at the edge of the remnant's rim and therefore can be approximated as a slab moving uniformly across the sky. This would not be the case if there was significant curvature of the remnant shell in the line of sight, or if we were viewing through the center of the remnant where the back part of the shell would be moving away from us while the front part was moving toward us. So the proximity of the Cygnus Loop to the Earth again favors resonance line scattering. As for (c), very little information is available. One location where turbulence likely occurs is at the contact discontinuity of the ejecta in young SNRs. The instabilities that operate here drive turbulent eddies and vortices that imprint a random pattern of velocity flows onto the overall radial expansion of the ejecta. Any such turbulence has an effect of increasing the root-mean-square velocity broadening $v$ resulting in a reduction of the scattering cross section. The forward shock in the adiabatic phase of remnant evolution is hydrodynamically stable and will not suffer this effect. But radiative cooling at the forward shock causes hydrodynamical instabilities (Blondin et al. 1998). Ultimately the amount of line broadening due to turbulence of this type will be some fraction of the shock velocity, so that younger SNRs will tend to have a higher line broadening and hence lower cross-section for resonance line scattering. If the part of the Cygnus Loop we observed is still in the Sedov phase of evolution, the conditions for minimal turbulence are met. Therefore, according to all three items discussed above, the Cygnus Loop is favored to have a significant effect due to resonance line scattering. Since the Cygnus Loop appears to be a rim-brightened shell in the X-ray band, optical depth effects may only play an important role at the rim where the plasma path lengths are largest. The abundance of O relative to other heavy elements has been show to decrease at the shell region of the Cygnus Loop (e.g. Miyata \\& Tsunemi 1999). Likewise, Leahy (2004) analyzed Chandra data of the bright southwestern region and found that the abundances of O group elements (C, N, and O) is roughly half those of the Ne group (Ne, Na, Mg, Al, Si, S, and Ar) and Fe group (Ca, Fe, and Ni). The depleted O abundance relative to other elements is consistent with our results and suggests that optical depth effects might be important at the southwestern region too. It should be noted that Miyata \\& Tsunemi (1999) and Leahy (2004) assumed the C and N abundances relative to O to be solar because they had neither high enough spectral resolution nor sufficient effective area for detecting C and N lines. The Suzaku XIS camera enables us to determine the abundances of C, N, and O independently for diffuse X-ray sources so future additional observations with Suzaku are essential for abundance measurements. Katsuda et al.~(2007) analyzed four Suzaku pointings of the northeastern region of the Cygnus Loop and showed that O is relatively depleted by a factor of two compared with other heavy elements. This result also supports our view for a significant optical depth effect. The Suzaku observatory has so far performed 8 mapping observations from the northeastern toward the southwestern region of the Cygnus Loop. Detailed analysis of the radial distribution of intensities of O$\\;${\\scriptsize\\rmfamily{VII}}\\,K$\\alpha$ and O$\\;${\\scriptsize\\rmfamily{VIII}}\\,K$\\alpha$ as well as their intensity ratios may help to clarify optical depth effects in the SNR. However, for this to work it will be necessary to separate out the effects of any intrinsic radial variation in the plasma temperature, which can also produce variations in the observed line intensity ratio. In clusters of galaxies, an anomalous ratio of Fe K$\\beta$ to K$\\alpha$ emission lines has been observed, which is offered as evidence for resonance line scattering in the relatively dense central cluster plasma (e.g., Tawara 1995). Unfortunately, due to their NEI condition, the intensity ratio of the K$\\beta$ to K$\\alpha$ emission line complexes is not a good indicator of resonance line scattering in SNRs. For definitive answers, we need to resolve individual K-shell lines of He-like ions as well as the Ly$\\alpha$ and Ly$\\beta$ lines for a range of elemental species as a function of position across the extent of diffuse thermal X-ray sources. Hopefully in the near future X-ray microcalorimeters onboard $XEUS$, Con-X, and/or $NeXT$ will provide this capability and allow X-ray astronomers to use resonance scattering and other techniques to gain a deeper insight into the nature of SNRs. \\vspace{1cm} We are grateful to all the other members of the Suzaku team. EM is supported by Grant-in-Aid for Specially Promoted Research (16002004) and the 21st Century COE Program, \\lq{\\it Towards a new basic science: depth and synthesis}\\rq. JPH acknowledges support from NASA grant NNG05GP87G. This research has made use of data obtained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center." }, "0801/0801.4768_arXiv.txt": { "abstract": "We present results on the interstellar medium (ISM) properties of 29 galaxies based on a comparison of {\\it Spitzer} far-infrared and Westerbork Synthesis Radio Telescope radio continuum imagery. Of these 29 galaxies, 18 are close enough to resolve at $\\la$1~kpc scales at 70~$\\micron$ and 22~cm. We extend the \\citet{ejm06a,ejm06b} approach of smoothing infrared images to approximate cosmic-ray (CR) electron spreading and thus largely reproduce the appearance of radio images. Using a wavelet analysis we decompose each 70~$\\micron$ image into one component containing the star-forming {\\it structures} and a second one for the diffuse {\\it disk}. The components are smoothed separately, and their combination compared to a free-free corrected 22~cm radio image; the scale-lengths are then varied to best match the radio and smoothed infrared images. We find that late-type spirals having high amounts of ongoing star formation benefit most from the two-component method. We also find that the disk component dominates for galaxies having low star formation activity, whereas the structure component dominates at high star formation activity. We propose that this result arises from an age effect rather than from differences in CR electron diffusion due to varying ISM parameters. The bulk of the CR electron population in actively star-forming galaxies is significantly younger than that in less active galaxies due to recent episodes of enhanced star formation; these galaxies are observed within $\\sim$10$^{8}$~yr since the onset of the most recent star formation episode. The sample irregulars have anomalously low best-fit scale-lengths for their surface brightnesses compared to the rest of the sample spirals which we attribute to enhanced CR electron escape. ", "introduction": "The interstellar medium (ISM) is a complex environment comprised of a diverse mix of extremely tenuous matter (by terrestrial standards) spanning a range of energetic states. Atomic, ionized, and molecular material make up a series of gaseous ({\\it thermal}) phases usually categorized by their temperature and density; combinations of these components are used to define the so-called two- and three-phase models of the ISM and variations thereof \\citep[e.g.][and references therein]{cm95}, however, this is not the entire picture. There is an additional phase of the ISM which is often overlooked due to the difficulties associated with making direct observations of its constituents. This is the relativistic ({\\it non-thermal}) phase which is made up of relativistic charged particles known as cosmic rays (CRs) and magnetic fields. The CRs within galaxies have an energy density comparable to that of the gaseous phases \\citep[e.g.][]{bc90}. They fill up the entire volume of galaxies and are important sources of heating and ionization of the ISM, though characterizing their propagation remains an ongoing astrophysical problem \\citep[see][for a review]{smp07}. The relativistic components of the ISM are important dynamically and may play a significant role in the regulation of star formation during the formation and evolution of galaxies \\citep[e.g.][and references therein]{kf01,cox05}. Magnetic fields both help to support interstellar matter against a galaxy's gravitational potential and confine CRs to galaxy disks; thus, magnetic fields and CRs take part in the hydrostatic balance and stability of the ISM while possibly determining the properties of gas spiral arms \\citep{beck07} and even aiding in the triggering of star formation \\citep{msw74,be82}. For sufficiently large CR pressures Parker instabilities \\citep{ep66} can create breaches in magnetic disks, allowing CRs and interstellar material to freely stream into intergalactic space. To date, most of our knowledge about the relativistic ISM outside of the Galaxy has been obtained indirectly through the detection of synchrotron emission via multi-frequency radio observations \\citep[e.g.][]{nd91,dlg95,ul96,ji99,rb05}. Synchrotron emission arises from CR electron energy losses as these particles are accelerated in the magnetic fields of galaxies. Although the energy density in CR electrons is only $\\sim$1\\% of that for CR nuclei, the similarity between the spatial distributions of gamma-ray and synchrotron emission within the Galaxy suggests that CR electrons and CR nuclei are fairly well mixed on the scales of a few hundred parsecs \\citep[e.g.][]{ch82,jb86,ww91}. The spatial distribution of a galaxy's synchrotron emission is a function of a galaxy's CR electron and magnetic field distributions. Thus, radio synchrotron maps provide only limited insight on the source distribution of the CR electrons as well as the distances the particles may have traveled before ending up in their current location of emission. Massive stars ($\\ga$8~$M_{\\sun}$) are the progenitors of supernovae (SNe) whose remnants (SNRs), through the process of diffusive shock acceleration \\citep{ab78,bo78}, appear to be the main acceleration sites of CR electrons responsible for a galaxy's observed synchrotron emission. These same young massive stars are often the primary sources for dust heating as they emit photons which are re-radiated at far-infrared (FIR) wavelengths. This shared origin between the FIR and radio emission of galaxies is thought to be the foundation for the observed FIR-radio correlation among \\citep[e.g.][]{de85,gxh85,sn97,nb97,yun01} and within galaxies \\citep[e.g.][]{bg88,Xu92,mh95,hoer98,hip03,ejm06a,ah06}. \\begin{deluxetable*}{ccccccccccccc} \\tablecaption{Basic Galaxy Data\\label{tbl-galdat}} \\tablewidth{0pt} \\tabletypesize{\\scriptsize} \\tablehead{ \\colhead{} & \\colhead{R.A.} & \\colhead{Decl.} & \\colhead{$D_{25}$} & \\colhead{} & \\colhead{} & \\colhead{$M_{\\rm opt}$} & \\colhead{$W_{20}$/$W_{20}^{0}$} & \\colhead{$V_{r}$} & \\colhead{Dist.} & \\colhead{$i$} & \\colhead{PA} & \\colhead{Distance}\\\\ \\colhead{Galaxy} & \\colhead{(J2000)} & \\colhead{(J2000)} & \\colhead{(arcmin)} & \\colhead{Type} & \\colhead{Nuc.} & \\colhead{(mag)} & \\colhead{(km s$^{-1}$)} & \\colhead{(km s$^{-1}$)} & \\colhead{(Mpc)} & \\colhead{($\\degr$)} & \\colhead{($\\degr$)} & \\colhead{References}\\\\ \\colhead{(1)} & \\colhead{(2)} & \\colhead{(3)} & \\colhead{(4)} & \\colhead{(5)} & \\colhead{(6)} & \\colhead{(7)} & \\colhead{(8)} & \\colhead{(9)} & \\colhead{(10)}& \\colhead{(11)}& \\colhead{(12)}& \\colhead{(13)} } \\startdata NGC~628 & 1 36 41.7 & +15 46 59 & 10.5$\\times$9.5 & SAc & \\ldots\t & -20.9\t& 74\t /175 & 657 & 7.3 &25 & 25 & 1 \\\\ NGC~925 & 2 27 17.0 & +33 34 43 & 10.5$\\times$5.9 & SABcd & H~\\textsc{II} & -20.6\t& 224\t /267 & 553 & 9.1 &57 &102 & 2 \\\\ NGC~2403 & 7 36 51.4 & +65 36 09 & 21.9$\\times$12.3& SABcd & H~\\textsc{II} & -19.7\t& 257\t /306 & 131 & 3.2 &57 &127 & 2 \\\\ Holmb~II & 8 19 04.0 & +70 43 09 & 7.9$\\times$6.3 & Im & \\ldots\t & -17.1\t& 73\t /121 & 157 & 3.4 &37 & 15 & 3 \\\\ NGC~2841 & 9 22 02.6 & +50 58 35 & 8.1$\\times$3.5 & SAb & Lin/Sy1\t & -20.7\t& 611\t /664 & 638 &14.1 &67 &147 & 4 \\\\ NGC~2976 & 9 47 15.3 & +67 55 00 & 5.9$\\times$2.7 & SAc & H~\\textsc{II} & -17.6\t& \\ldots/\\ldots & 3 & 3.6 &64 &143 & 3 \\\\ NGC~3031 & 9 55 33.2 & +69 03 55 & 26.9$\\times$14.1& SAab & Lin\t & -21.2\t& 446\t /515 & -34 & 3.6 &60 &157 & 2 \\\\ NGC~3184 &10 18 16.9 & +41 25 28 & 7.4$\\times$6.9 & SABcd & H~\\textsc{II} & -19.0\t& 142\t /396 & 592 &11.1 &21 &135 & 5 \\\\ NGC~3198 &10 19 54.9 & +45 32 59 & 8.5$\\times$3.3 & SBc & \\ldots\t & -20.2\t& 318\t /343 & 663 &13.7 &68 & 35 & 2 \\\\ IC~2574 &10 28 21.2 & +68 24 43 & 13.2$\\times$5.4 & SABm & \\ldots\t & -17.7\t& 123\t /134 & 57 & 4.0 &67 & 50 & 6 \\\\ NGC~3627 &11 20 15.0 & +12 59 30 & 9.1$\\times$4.2 & SABb & Sy2\t & -20.8\t& 378\t /417 & 727 & 9.4 &65 &173 & 2 \\\\ NGC~3938 &11 52 49.5 & +44 07 14 & 5.4$\\times$4.9 & SAc & \\ldots\t & -20.1\t& 112\t /265 & 809 &13.3 &25 & 0 & 7 \\\\ NGC~4125 &12 08 05.8 & +65 10 27 & 5.8$\\times$3.2 & E6p & \\ldots\t & -21.6\t& \\ldots/\\ldots &1356 &22.9 &58 & 95 & 8 \\\\ NGC~4236 &12 16 42.1 & +69 27 46 & 21.9$\\times$7.2 & SBdm & \\ldots\t & -18.1\t& 176\t /185 & 0 & 4.5 &72 &162 & 3 \\\\ NGC~4254 &12 18 49.5 & +14 24 59 & 5.4$\\times$4.7 & SAc & \\ldots\t & -21.6\t& 272\t /544 &2407 &16.6 &30 & 0 & 7$^{*}$\\\\ NGC~4321 &12 22 54.9 & +15 49 21 & 7.4$\\times$6.3 & SABbc & Lin\t & -22.1\t& 283\t /534 &1571 &14.3 &32 & 30 & 8 \\\\ NGC~4450 &12 28 29.5 & +17 05 06 & 5.2$\\times$3.9 & SAab & Lin\t & -21.4\t& 290\t /433 &1954 &16.6 &42 &175 & 7$^{*}$ \\\\ NGC~4552 &12 35 39.8 & +12 33 23 & 5.1$\\times$4.7 & E & \\ldots\t & -20.8\t& \\ldots/\\ldots & 340 &15.9 &23 & 0 & 9 \\\\ NGC~4559 &12 35 57.7 & +27 57 36 & 10.7$\\times$4.4 & SABcd & H~\\textsc{II} & -21.0\t& 251\t /273 & 816 &10.3 &67 &150 & 7 \\\\ NGC~4569 &12 36 49.8 & +13 09 46 & 9.5$\\times$4.4 & SABab & Lin/Sy\t & -22.0\t& 360\t /397 &-235 &16.6 &65 & 23 & 7$^{*}$\\\\ NGC~4631 &12 42 08.0 & +32 32 26 & 15.5$\\times$2.7 & SBd & \\ldots\t & -20.6\t& 320\t /322 & 606 & 7.7 &83 & 86 & 10 \\\\ NGC~4725 &12 50 26.6 & +25 30 06 & 10.7$\\times$7.6 & SABab & Sy2\t & -22.0\t& 410\t /570 &1206 &11.9 &46 & 35 & 2 \\\\ NGC~4736 &12 50 53.0 & +41 07 14 & 11.2$\\times$9.1 & SAab & Lin\t & -19.9\t& 241\t /400 & 308 & 5.0 &37 &105 & 8 \\\\ NGC~4826 &12 56 43.7 & +21 40 52 & 10.0$\\times$5.4 & SAab & Sy2\t & -20.3\t& 311\t /363 & 408 & 5.0 &59 &115 & 7 \\\\ NGC~5033 &13 13 27.5 & +36 35 38 & 10.7$\\times$5.0 & SAc & Sy2\t & -20.9\t& 446\t /501 & 875 &14.8 &63 &170 & 7 \\\\ NGC~5055 &13 15 49.2 & +42 01 49 & 12.6$\\times$7.2 & SAbc & H~\\textsc{II}/Lin & -19.0\t& 405\t /489 & 504 & 7.8 &56 &105 & 7 \\\\ NGC~5194 &13 29 52.7 & +47 11 43 & 11.2$\\times$6.9 & SABbc & H~\\textsc{II}/Sy2 & -21.4\t& 195\t /244 & 463 & 7.8 &53 &163 & 7 \\\\ NGC~6946 &20 34 52.3 & +60 09 14 & 11.5$\\times$9.8 & SABcd & H~\\textsc{II} & -21.3\t& 242\t /457 & 48 & 6.8 &32 & 69 & 11 \\\\ NGC~7331 &22 37 04.1 & +34 24 56 & 10.5$\\times$3.7 & SAb & Lin & -21.8\t& 530\t /561 & 816 &14.5 &71 &171 & 2 \\enddata \\tablecomments{Col. (1): ID. Col. (2): The right ascension in the J2000.0 epoch. Col. (3): The declination in the J2000.0 epoch. Col. (4): Major- and minor-axis diameters. Col. (5): RC3 type. Col. (6): Nuclear type: H~\\textsc{II}: H~\\textsc{II} region; Lin: LINER; Sy: Seyfert (1, 2). Col. (7): Absolute R magnitude, when available; otherwise from the V or B bands. Col. (8): Observed/inclination-corrected 21 cm neutral hydrogen line width at 20\\% of maximum intensity, in km~s$^{-1}$, taken from \\citet{tul88} or RC3. Col. (9): Heliocentric velocity. Col. (10): Distance in Mpc Col. (11): Inclination in degrees. Col. (12): Position Angle in degrees. (13): Distance References} \\tablerefs{(1)~\\citet{sfe01}; (2)~\\citet{wf01}; (3)~\\citet{ik02}; (4)~\\citet{lm01}; (5)~\\citet{dl02}; (6)~\\citet{ik03}; (7)~K. Masters 2007, in preparation: ($^{*})$ indicates distance set to Virgo Cluster center; (8)~\\citet{jt01}; (9)~\\citet{lf00}; (10)~\\citet{as05}; (11)~\\citet{ik00}} \\end{deluxetable*} Coupling the shared origin of a galaxy's FIR and radio emission with the fact that the mean free path of dust-heating photons ($\\sim$100~pc) is significantly shorter than the expected diffusion length of CR electrons ($\\sim$1-2~kpc) led \\citet{bh90} to conjecture that the radio image of a galaxy should resemble a smoothed version of its infrared image. Consequently, it appears that the close spatial correlation between the FIR and radio continuum emission within galaxies can be used to characterize the propagation history of CR electrons. This prescription has been shown to hold for galaxies observed at the ``super resolution'' ($\\la$1$\\arcmin$) of {\\it IRAS} HIRES data \\citep{mh98} and, more recently, for high resolution ($\\sim$18$\\arcsec$) {\\it Spitzer} 70$~\\micron$ imaging \\citep[][hereafter M06a]{ejm06a}. This phenomenology has been further corroborated on scales $\\ga$50~pc by \\citet{ah06} who find synchrotron haloes around individual star-forming regions are more extended than FIR-emitting regions within the Large Magellanic Cloud. \\citet[][hereafter, M06b]{ejm06b} recently studied how the spatial distributions of a galaxy's FIR and radio emission vary as a function of the intensity of star formation. They concluded that CR electrons are, on average, younger and closer to their place of origin within galaxies having higher amounts of star formation activity compared with more quiescent galaxies. Using a wavelet-based image decomposition, we extend this work by attempting to characterize separately CR electron populations associated with a galaxy's diffuse disk and its star-forming complexes. We carry out this study for a sample of galaxies observed as part of the {\\it Spitzer} Infrared Nearby Galaxies Survey \\citep[SINGS;][]{rk03} and the Westerbork Synthesis Radio Telescope \\citep[WSRT-SINGS;][]{rb07} for which we have the spatial resolution to resolve physical scales $<1$~kpc. The paper is organized as follows: The galaxy sample is defined in $\\S$2 while the observations and data reduction techniques are discussed in $\\S$3. In $\\S$4 we present and discuss a correlation analysis between FIR/radio ratios and various other physical quantities on sub-kiloparsec scales within galaxies; this section leads us to confirm quantitatively that our favored phenomenological model is the best description for the FIR-radio correlation within galaxies. In $\\S$5 we describe our wavelet-based, two-component image-smearing analysis; the corresponding results are then presented in $\\S$6 and their physical implications are discussed in $\\S$7. We briefly discuss outstanding issues and future prospects in $\\S$8 and summarize our results and conclusions in $\\S$9. ", "conclusions": "} Using a two-component image-smearing analysis, we have separated the signatures of CR electron diffusion at spatial scales corresponding to star-forming structures ($<$1~kpc) and galaxy disks ($\\geq$1~kpc) within 18 galaxies observed as part of SINGS and WSRT-SINGS. Our results and conclusions can be summarized as follows: \\begin{enumerate} \\item We confirm and extend earlier results of M06a,b. Empirically, the dispersion in the FIR-radio correlation within galaxies is most reduced by an image-smearing model; this improvement is significantly better than what can be achieved by fitting correlations and removing linear trends. \\item The best-fit global scale-lengths decrease as a function of increasing star formation activity as measured by the infrared surface brightness of a galaxy. Our interpretation is that a galaxy's CR electrons are closer to their place of origin within galaxies having intense star formation activity. \\item The trend of decreasing best-fit {\\it global} scale-length with increasing radiation field energy density is due to higher surface brightness galaxies having undergone a recent enhancement of star formation rather than variations in other ISM parameters. For sufficiently large enhancements, these galaxies are observed within $\\sim$10$^{8}$~yr of the onset of the most recent star formation episode. \\item Unlike spirals, irregular galaxies lack any well defined diffuse disk component at either 70~$\\micron$, or especially at 22~cm. Presumably, the CR electrons escape these galaxies soon after leaving their parent star-forming regions due the absence of a dense ISM which would keep large-scale interstellar magnetic field locked into place. This conclusion helps to explain why these galaxies have global FIR/radio ratios systematically greater than canonical values. \\item As infrared surface brightness increases, the characteristic diffusion scale-length of a galaxy's CR electron population begins to transition at \\(\\log~U_{\\rm rad} \\leq -12.5\\), or \\(\\log~\\Sigma_{\\rm SFR} \\leq -2.3\\), from being biased by CR electrons making up its diffuse disk to being biased by those recently injected near star-forming structures. From this we conclude that a galaxy's CR electron population transitions from being dominated by old CR electrons to being dominated by young CR electrons as a function of star formation intensity. \\item The two-component analysis works better than smearing with a single smoothing kernel for spiral galaxies of type Sb or later which have high amounts of ongoing star formation activity (i.e. $\\sim$40\\% of the sample). This result suggests that star formation must be intense and highly structured for the two-component analysis of these data to differentiate properly between the different CR electron populations. \\end{enumerate}" }, "0801/0801.1651_arXiv.txt": { "abstract": " ", "introduction": "Many studies have showcased the great potential of the LHC for producing and discovering supersymmetric particles \\cite{lhc,cmstdr,LHC}, and the ability of experiments at a linear $e^+ e^-$ collider (LC) to measure sparticle properties in detail, if their pair-production thresholds lie within its kinematic reach \\cite{ilc}. Most of these studies have assumed that $R$ parity is conserved, in which case the lightest supersymmetric particle (LSP) may provide the cold dark matter postulated by astrophysicists and cosmologists \\cite{EHNOS}. Further, most studies have been within the framework of the minimal supersymmetric extension of the Standard Model (MSSM) \\cite{mssm}, and assumed that the LSP is the lightest neutralino $\\chi$. We also adopt this framework in this paper. In this case, the classic signature of sparticle pair production is missing energy carried away by the dark matter particles $\\chi$. Studies have indicated that experiments at the LHC should be able to detect gluinos and squarks weighing up to $\\sim 2.5$~TeV \\cite{Baer}, whereas any sparticles weighing less than the beam energy should be detectable at a LC. One specific supersymmetric version of this framework that has commonly been examined is the Constrained MSSM (CMSSM)~\\cite{funnel,cmssm,efgosi,eoss,cmssmwmap}, in which the soft supersymmetry-breaking mass parameters are assumed to be universal at some high scale, generally taken to be the supersymmetric GUT scale, $M_{GUT} \\sim 10^{16}$ GeV. Within the CMSSM, renormalization group equations (RGEs) can be used to calculate the weak-scale observables in terms of four continuous and one discrete parameter; the scalar mass, $m_0$, the gaugino mass, $m_{1/2}$, and the trilinear soft breaking parameter, $A_0$ (each specified at the universality scale), as well as the ratio of the Higgs vevs, $\\tbt$, and the sign of the Higgs mixing parameter, $\\mu$. The reaches of colliders such as the LHC or a LC are then often expressed in the $(m_{1/2}, m_0)$ plane for representative values of $A_0, \\tbt$ and the sign of $\\mu$. However, the mechanism of supersymmetry breaking is not known, and alternative scenarios should also be considered. Rather than postulate that the soft supersymmetry-breaking parameters are universal at some GUT scale, one might consider theories in which this universality assumption for the the soft supersymmetry-breaking parameters is relaxed. One possibility, motivated to some extent by supersymmetric GUT scenarios and the absence of flavour-changing interactions due to sparticle exchanges, would be to relax (for example) the universality assumption for the soft supersymmetry-breaking contributions to the Higgs scalar masses at the GUT scale (the NUHM) \\cite{nonu,nuhm}, and more radical abandonments of universality could also be considered. We consider here a different generalization of the CMSSM, in which universality of the soft supersymmetry-breaking mass parameters is maintained, but is imposed at some lower input scale $M_{in} < M_{GUT}$ \\cite{eos1,eos2}. Such GUT-less (or sub-GUT) scenarios may arise in models where the dynamics that breaks or communicates supersymmetry breaking to the observable sector has an intrinsic scale below $M_{GUT}$, and switches off at higher scales, much as the effective dynamical quark mass in QCD switches off at scales $> \\Lambda_{QCD}$. Mirage unification scenarios~\\cite{mixed} offer one class of examples in which the low-energy evolution of the gaugino masses is as if they unified at some scale $< M_{GUT}$. In principle, one could consider scenarios in which universality is imposed on the different MSSM soft supersymmetry breaking parameters $m_{1/2}, m_0$ and $A_0$ at different input scales $M_{in}$. However, here we follow~\\cite{eos1,eos2} in studying the simplest class of GUT-less scenarios with identical $M_{in}$ for all the soft supersymmetry-breaking parameters. As one would expect, the reduction in the universality scale has important consequences for the low-energy sparticle mass spectrum. In particular, the hierarchy of gaugino masses familiar in the GUT-scale CMSSM is reduced with, for example, a substantial reduction in the ratio of gluino and bino masses. Likewise, squark and slepton masses also approach each other as $M_{in}$ is reduced. These effects have important consequences for the $(m_{1/2}, m_0)$ planes in GUT-less scenarios: for example, the boundaries imposed by the absence of a charged ${\\tilde \\tau_1}$ LSP and the generation of an electroweak symmetry breaking vacuum approach each other as $M_{in}$ decreases. A corollary of the `squeezing' of the sparticle mass spectrum is the observation made in \\cite{eos1} and \\cite{eos2} that, as the universality scale $M_{in}$ is decreased from the GUT scale, there are dramatic changes in the cosmological constraint imposed on the parameter space by the relic density of neutralinos inferred from WMAP and other observations~\\cite{WMAP}. In general, as $M_{in}$ decreases, the regions where the relic neutralino LSP density falls within the range preferred by WMAP and other measurements~\\cite{WMAP} tend to move to larger $m_{1/2}$ and $m_0$. This implies that, whereas in the GUT-scale CMSSM the relic neutralino is {\\it overdense} in most of the region with $m_{1/2}, m_0 < 1$~TeV, as $M_{in}$ decreases to $\\sim 10^{11}$~GeV most of this region becomes {\\it underdense}. In this paper, we consider the implications of these observations for the prospects for sparticle detection at the LHC and a LC. ATLAS and CMS have estimated their reaches in inclusive supersymmetry searches for multiple jets and missing transverse energy, as functions of the accumulated and analyzed LHC luminosity, which may be expressed as reaches for gluino and squark masses~\\cite{cmstdr}. These may in turn be converted into the reaches in the $(m_{1/2}, m_0)$ planes for different values of $M_{in}$. The masses of weakly-interacting sparticles such as sleptons, charginos and neutralinos are determined across these $(m_{1/2}, m_0)$ planes, and hence the ATLAS/CMS reaches may be converted into the corresponding sparticle pair-production thresholds at a generic LC. These converted reaches may be interpreted in at least two ways. If the LHC {\\it does discover} supersymmetry, then one may estimate, within the CMSSM or any given GUT-less model, the {\\it maximum} centre-of-mass energy that would suffice for a LC to make detailed follow-up measurements of at least some sparticles. Conversely, if the LHC {\\it does not discover} supersymmetry within a given physics reach, one can, within the CMSSM or any given GUT-less model, estimate the {\\it minimum} centre-of-mass energy below which a LC would not provide access to any sparticles. In general, because of the `squeezing' of the sparticle mass spectrum as $M_{in}$ {\\it decreases}, for any given LHC physics reach the required LC centre-of-mass energy {\\it increases} correspondingly. This argument can be carried through whether one disregards the cosmological density of dark matter entirely, or regards it solely as an upper limit on the relic LSP density, or interprets it as a narrow preferred band. In the third case, the prospects for sparticle detection at the LHC recede with the preferred dark matter regions in the $(m_{1/2}, m_0)$ planes as $M_{in}$ decreases. Within the specific preferred dark-matter regions, the relation between the LHC and LC reaches can be made more precise. For example, if the LHC discovers sparticles with 1~fb$^{-1}$ of data, within the CMSSM a centre-of-mass energy of 600~GeV would suffice for a LC to to produce pairs of neutralinos, if they provide the cold dark matter, whereas over 1 TeV might be required in a GUT-less model with $M_{in} > 10^{11.5}$~GeV. These required energies increase to 800~GeV in the CMSSM and 1.4~TeV in GUT-less models with $M_{in} > 10^{11.5}$~GeV if the LHC requires 10~fb$^{-1}$ to discover supersymmetry. ", "conclusions": "We have discussed in the previous Section how much centre-of-mass energy would be required to `guarantee' the observability of sparticle pair production in $e^+ e^-$ collisions under various hypotheses for the integrated luminosity required for discovering supersymmetry at the LHC and for different values of the universality scale $M_{in}$. We have also discussed how corresponding sparticle exclusions at the LHC would set lower limits on the possible thresholds for producing different sparticle pairs at a LC. To conclude, we now consider the capabilities of LCs with various specific proposed centre-of-mass energies. Even if supersymmetry were to be found at the LHC with 1~fb$^{-1}$ of integrated luminosity, a LC with $E_{cm} = 0.5$~TeV would not be `guaranteed' to produce $\\chi \\chi$ pairs or other sparticle pairs. However, even if supersymmetry were to be excluded in the LHC's 1~fb$^{-1}$ discovery region, the possibility of observing sparticles at a LC with $E_{cm} = 0.5$~TeV could not be excluded for $M_{in} > 10^{13.5}$~GeV, and such a LC might also pair-produce charged sparticles if $M_{in} > 10^{15}$~GeV and/or produce $\\chi \\chi_2$ in association if $M_{in} > 10^{14.5}$~GeV. On the other hand, if supersymmetry were not even within the 10~fb$^{-1}$ discovery reach of the LHC, a LC with $E_{cm} = 0.5$~TeV might be (barely) above the $\\chi \\chi$ threshold only if $M_{in} \\gtrsim 10^{15.5}$~GeV, and there would be no likelihood of charged-sparticle or $\\chi \\chi_2$ production. A LC with $E_{cm} = 1$~TeV would be `guaranteed', if supersymmetry were to be found at the LHC with 1~fb$^{-1}$ of integrated luminosity, to produce $\\chi \\chi$ pairs in any GUT-less scenario with $M_{in} > 10^{12}$~GeV. Analogous `guarantees' for charged-sparticle pair production or associated $\\chi \\chi_2$ production could be given only for $M_{in} > 10^{13} (10^{14})$~GeV, respectively. On the other hand, if supersymmetry were not even within the 10~fb$^{-1}$ discovery reach of the LHC, it might still be possible to find $\\chi \\chi$ (charged-sparticle pairs) ($\\chi \\chi_2$) at a LC if $M_{in} > 10^{12.5} (10^{13.3}) (10^{13})$~GeV. Finally, even if the LHC would require 10~fb$^{-1}$ to discover supersymmetry, a LC with $E_{cm} = 1.5$~TeV would be `guaranteed' to produce $\\chi \\chi$ and $\\chi \\chi_2$ pairs in all the allowed WMAP-compatible scenarios, and charged-sparticle pair production would be `guaranteed' for all except a small range of $M_{in}$ between $10^{12}$ and $10^{13}$~GeV. Hence, a LC with $E_{cm} = 1.5$~TeV would be well matched to the physics reach of the LHC with this luminosity, whereas a LC with a lower $E_{cm}$ might well be unable to follow up on a discovery of supersymmetry at the LHC. However, as already mentioned, even in the absence of any `guarantee', it could still be that the LHC discovers supersymmetry at some mass scale well below the limit of its sensitivity with 10~fb$^{-1}$ of integrated luminosity, in which case a lower-energy LC might still have interesting capabilities to follow up on a discovery of supersymmetry at the LHC. It is clear that the physics discoveries of the LHC will be crucial for the scientific prospects of any future LC. Supersymmetry is just one of the scenarios whose prospects at a LC may depend on what is found at the LHC. Even within the supersymmetric framework, there are many variants that should be considered. Even if $R$ parity is conserved, the LSP might not be the lightest neutralino. Even if it is, the relevant supersymmetric model may not be minimal. Even if it is the MSSM, supersymmetry breaking may not be universal. Even if it is, the universality scale may not be the same for gauginos and sfermions. Nevertheless, we hope that study serves a useful purpose in highlighting some of the issues that may arise in guessing the LC physics prospects on the basis of LHC physics results." }, "0801/0801.3462_arXiv.txt": { "abstract": "We consider a magnetic Bianchi I braneworld, embedded in between two Schwarzschild-AdS spacetimes, boosted equal amounts in opposite directions and compare them to the analagous solution in four-dimensional General Relativity. The efficient dissipation of anisotropy on the brane is explicitly demonstrated, a process we dub braneworld isotropization. From the bulk point of view, we attribute this to anisotropic energy being carried into the bulk by hot gravitons leaving the brane. From the brane point of view this can be interpreted in terms of the production of particles in the dual CFT. We explain how this result enables us to gain a better understanding of the behaviour of anisotropic branes already studied in the literature. We also show how there is evidence of particles being over-produced, and comment on how this may ultimately provide a possible observational signature of braneworlds. ", "introduction": "The most recent data from WMAP suggests that the Cosmic Microwave Background (CMB) is isotropic to within one part in $10^5$~\\cite{wmap}. Why is our universe {\\it so} isotropic? This has been an important question in cosmology ever since the CMB's discovery back in the mid 1960s~\\cite{cmb}. The obvious way to approach this is to assume that the universe began in a highly {\\it an}isotropic state, and ask what dynamical mechanism caused the universe to shed nearly all its anisotropy, leaving behind the highly symmetric state we observe today. The most popular mechanism is, of course, inflation~\\cite{inflation}. In inflation, almost all anisotropy was diluted away during a period of accelerated expansion in the early universe. The accelerated expansion was driven by a scalar field rolling slowly down its potential, closely resembling a positive cosmological constant. Indeed, it can be shown that in the presence of a positive cosmological constant, all but one of the Bianchi models\\footnote{The exception is Bianchi IX.}, describing a homogeneous but anisotropic cosmology, evolve asymptotically to an isotropic de Sitter solution~\\cite{wald}. However, there is still an ongoing discussion on how fine-tunned the initial conditions should be to get a sufficiently flat, homogeneous and isotropic universe after inflation (see \\cite{carrol} and references therein). It is often easy to forget that before inflation, plenty of other ``isotropization''mechanisms were put forward, with varying degrees of success. The earliest was probably the ``phoenix'', or ``oscillatory'', universe~\\cite{phoenix}, which expands for a while, recollapses towards a bounce, and then starts to expand again. As entropy builds up after each oscillation, it was argued that the period of oscillation increased, and the universe approached an isotropic, homogeneous state of thermal equilibrium. Although these ideas ran into problems with inevitable singularities~\\cite{sings}, they have been revived recently with the advent of the cyclic universe~\\cite{cyclic} and a possible resolution to its singularity~\\cite{niz}. Another means of damping anisotropy is particle production due to quantum effects at very early times~\\cite{pp}. Energetically, this can be understood as anisotropy being converted into thermal radiation. If, for example, the universe were contracting along one direction, but expanding along the others, the momenta of virtual particles would be blueshifted along the contracting direction. This would make the virtual particles more likely to materialize as real particles, drawing energy from the contraction, and lowering the amount of anisotropy. However, if we assume an arbitrarily large amount of initial anisotropy, a huge amount of thermal radiation would need to be produced to account for the levels of isotropy seen in the CMB, and as a result, the photon-baryon ratio would massively exceed its observed value~\\cite{bm}. Particle production alone cannot, therefore, account for enough dissipation, and we typically assume it to be one of many contributing factors, including neutrino viscosity~\\cite{viscosity}, and, inevitably, inflation. In this paper, we point out another mechanism for dissipating anisotropy: {\\it braneworld isotropization}. Consider a Randall-Sundrum type braneworld~\\cite{rs}, and assume it is highly anisotropic. As the brane evolves, the anisotropic energy on the brane leaks into the bulk, and the brane becomes rapidly more isotropic, whilst the bulk becomes { less} homogeneous and/or { less} isotropic. To get a feel for how this works consider the projection of the Einstein equation onto a Randall-Sundrum braneworld~\\cite{sms}, schematically given by \\be G_\\mn(\\gamma)=8 \\pi G_4 T_\\mn+G_5^2(T_\\mn)^2-E_\\mn, \\ee where $G_4$ and $G_5$ are the brane and bulk Newton's constants respectively, and $T_\\mn$ is the energy-momentum tensor for matter on the brane. Note that we have a non-local piece, $E_\\mn$, which is the electric part of the bulk Weyl tensor, projected onto the brane. The ``anisotropic energy density'' on the brane, in the sense described in~\\cite{viscosity}, is stored in the local geometrical piece, $G_\\mn$, and as the brane evolves, this energy is transferred into the bulk through the bulk Weyl term, $E_\\mn$. The contribution of the bulk Weyl term is often best understood holographically (see, for example~\\cite{bwholo3}) using the braneworld version of the AdS/CFT correspondence~\\cite{adscft, bwholo1}. This states that {\\it Randall-Sundrum braneworld gravity is dual to a CFT, with a UV cut-off, coupled to gravity in $3+1$ dimensions}. It follows that a {\\it classical} source on the brane behaves like a {\\it quantum} source in $3+1$ dimensons, in the sense that it has been dressed with quantum corrections from the CFT~\\cite{bwholo2}. The anisotropy dissipation process we have just described can now be interpreted as particle production in the CFT. Energy is drawn from the anisotropy to fuel the production of CFT particles, leading to isotropization on the brane. There is already evidence in the literature for braneworld isotropization. Consider, for example, the quest to find anisotropic geometries supporting a perfect fluid. This is easy enough in four-dimensional General Relativity, the Bianchi or Kantowski-Sachs metrics being good examples~\\cite{kantowski, exacsols}. In brane cosmology, full solutions have not been so easy to find~\\cite{maar1, orsaylot, frolov}, and only a rather contrived set of kasner-like solutions are known exactly~\\cite{orsaylot, frolov}. Indeed, one can prove that if the bulk is static, an anisotropic brane cannot support a perfect fluid~\\cite{orsaylot}. This result can now be easily understood: particle production in the CFT leads to dissipation of anisotropy over time. From the $5D$ point of view, the leaking of anisotropy into the bulk prevents it from being static. Another interesting example of braneworld isotropization lies in the study of anisotropy dissipation in braneworld inflation~\\cite{maar2}. There it was noticed that inflation begins sooner than it would in $4D$ General Relativity. We attribute this to particle production in the CFT drawing its energy from the anisotopy, aswell as the kinetic energy of the inflaton. In this paper, we demonstrate braneworld isotropization explicitly by means of a concrete example. We generate the initial anisotropy on a Bianchi I braneworld by means of a constant magnetic field. Unlike much of the existing literature on anisotropic braneworlds the full bulk solution is known, corresponding to the planar limit of two boosted Schwarzschild-AdS black hole spacetimes, cut and pasted together to form the brane in the usual way. Because the brane embedding is highly non-trivial when viewed in global coordinates, it is convenient to work in a boosted coordinate system on either side of the brane, so that the bulk metric resembles the planar limit of two particular Myers-Perry-AdS spacetimes~\\cite{MP,HHT}. The boosts are equal and opposite on either side of the brane, so one may think of the corresponding Myers-Perry solutions as having equal and opposite angular momentum. We then track the brane's evolution, paying particular attention to the anisotropic shear, and compare it to the corresponding scenario with the same source in four-dimensional General Relativity. As expected, the shear anisotropy dissipates far quicker in the Randall-Sundrum braneworld scenario, than in $4D$ GR. A study of magnetic fields on a braneworld is also important in its own right. One of the great puzzles in astrophysics concerns the origin of large magnetic fields in galaxies and galaxy clusters (for an excellent review of magnetic fields in the early universe, see~\\cite{magrev}). It has been argued that these fields were generated by a galactic dynamo process~\\cite{dynamo}, sourced by a smaller pre-galactic ``seed'' field. However, one still has to account for the origin of the seed field, and even then the efficiency of this process has been brought into question~\\cite{dynamobad}. We are naturally led to consider the possibility that galatic fields result from large scale primordial fields left over from the big bang. In the standard cosmology, one can place bounds on the size of a large scale primordial magnetic field from nucleosynthesis ($B \\lesssim 10^{-7}$ G)~\\cite{BBNbound}, and from CMB temperature fluctuations ($B \\lesssim 10^{-9}$ G)~\\cite{CMBbound}. Given that the CMB bound is the stronger, we might speculate that this bound is weakened in a braneworld context, due to isotropization. Magnetic fields on branes have been considered in the past~\\cite{magbrane, ecc}, although a full knowledge of the bulk has been absent. Whilst it is true that one can certainly learn a lot {\\it without} full knowledge of the bulk, we believe it is a dangerous game to play. It is not at all obvious that a bulk solution that is regular near the brane evolves into something regular far from the brane. A good example of this is the black string solution~\\cite{bwbh} which is regular near the brane, but singular on the AdS horizon. The rest of this paper is organised as follows. In section~\\ref{sec:4dgr}, we review the Bianchi I solution for a constant magnetic field in four-dimensional General Relativity. For completeness, and in keeping with supernovae observations~\\cite{supernova}, we will allow for a positive cosmological constant that can be set to zero if necessary. In section~\\ref{sec:bw}, we introduce a certain Myers-Perry-AdS black hole in five dimensions, and take the planar limit, so that we end up with planar Schwarzschild-AdS in boosted coordinates. We then cut and paste two equal and oppositely boosted black holes onto one another, in the usual way, in order to form a Bianchi I braneworld, supported by brane tension, and a constant magnetic field. In section~\\ref{sec:compare} we compare the two scenarios, with special emphasis on the dissipation of anisotropy, and the effect of particle production in the CFT. Finally, in section~\\ref{sec:discuss}, we summarize our results. Our conventions follow the Landau-Lifshitz notation for the curvature tensors, and we use a $(-,+,...,+)$ signature for the metric. ", "conclusions": "\\lab{sec:discuss} In this paper, we have explicitly demonstrated the existence of a new phenomenon, called {\\it braneworld isotropization}, by means of a concrete example. We were able to embed a magnetic Bianchi I braneworld in between two Schwarzschild-$AdS_5$ spacetimes, boosted equal amounts in opposite directions. The magnetic field breaks isotropy on the brane, and leads to the production of thermal graviton radiation, that can carry energy-momentum into the bulk. The boosted coordinates mean that the brane observer sees equal and opposite momentum carried into the bulk by the hot gravitons leaving the brane. We can view braneworld isotropization as anisotropic energy on the brane being dumped into the bulk. From a completely $4D$ point of view, we can understand this effect using the AdS/CFT correspondence. In the Randall-Sundrum scenario, the gravitational physics in the bulk is equivalent to a strongly coupled CFT, cut-off in the UV, and minimally coupled to gravity on the brane. When isotropy is broken, the coupling of the CFT to gravity leads to particle production, the required energy being drawn from the anisotropy. This phenomenon can now be used to readily explain a number of existing results in the literature~\\cite{maar1, maar2, orsaylot, frolov}. For example, an anisotropic brane cannot support a perfect fluid in a static bulk~\\cite{orsaylot}, because the leaking of anisotropy off the brane prevents the bulk from being static, at least when viewed by an asymptotic braneworld observer. Note that in the solution described in this paper, we do not have a perfect fluid on the brane, and although the bulk is static according to an asymptotic observer using Schwarzschild time, it is {\\it not} static according to an asymptotic observer using the boosted time coordinate. It is the latter time coordinate that is used by the asymptotic braneworld observer at large $r(\\tau)$. It is worth noting that we have been able to prove that the solution presented here, along with the known isotropic solutions~\\cite{bcg}, are the most general solutions for a Bianchi I brane of the form (\\ref{metric4d}), embedded in Schwarzschild-$AdS_5$ with some form of ``sensible'' matter on the brane. We have not included the proof since it is lengthy, and not particularly illuminating. However, it does tell us that we will have to work a bit harder if we want to, say, include an arbitrary perfect fluid on the brane in addition to the magnetic field. This is a subject for future study. Nevertheless, we anticipate that many of the features seen here will remain, in particular, momentum being generated in the bulk by the magnetic field on the brane, and particle production in the CFT leading to anisotropy dissipation. Perhaps one of the most interesting features of this work was the possibility that one could detect a braneworld signature in the sky, as discussed near the end of section~\\ref{sec:bw}. In the braneworld picture, particles are actually {\\it over}-produced along the direction of the magnetic field, so that the expansion along that direction is {\\it slower} than in the orthogonal directions, in complete contrast with what happens in standard $4D$ General Relativity. For this reason it would be interesting to extend this work to other, more phenomenological, Bianchi models. If, say, one of the spatial directions were periodic, we might expect {\\it angular} momentum to be carried into the bulk, and so we would look for an embedding in the general Myers-Perry spacetimes~\\cite{MP, HHT}. A perturbative study along these lines has been considered~\\cite{Guth}, although a far more complete analysis is clearly required. In the long term, we would like to consider a physically realistic scenario with a Bianchi VIIh brane containing a homogeneous magnetic field along with an isotropic perfect fluid made up of matter, radiation and a cosmological constant. Of course, this would require an amalgamation of the ideas outlined in the previous two paragraphs. It would be very interesting to see what effect braneworld isotropization has on the CMB, and in particular the bounds on the size of the primordial magnetic field~\\cite{CMBbound}. We must also keep in mind constraints coming from nucleosynthesis. For example, if braneworld isotropization is {\\it too} efficient, then we might worry that CFT particle production will result in too much dark radiation, leading to an unacceptably low baryon density parameter." }, "0801/0801.2652_arXiv.txt": { "abstract": "{The distance to the Galactic Centre (GC) is of importance for the distance scale in the Universe. The value derived by Eisenhauer et al. (2005) of 7.62 $\\pm$ 0.32 kpc based on the orbit of one star around the central black hole is shorter than most other distance estimates based on a variety of different methods.} {To establish an independent distance to the GC with high accuracy. To this end Population-{\\sc ii} Cepheids are used that have been discovered in the \\OG-{\\sc ii} and \\OG-{\\sc iii} surveys.} {Thirty-nine Population-{\\sc ii} Cepheids have been monitored with the SOFI infrared camera on 4 nights spanning 14 days, obtaining typically between 5 and 11 epochs of data. Light curves have been fitted using the known periods from the OGLE data to determine the mean $K$-band magnitude with an accuracy of 0.01-0.02 mag. It so happens that 37 RR Lyrae stars are in the field-of-view of the observations and mean $K$-band magnitudes are derived for this sample as well.} {After correction for reddening, the period-luminosity relation of Population-{\\sc ii} Cepheids in the $K$-band is determined, and the derived slope of $-2.24 \\pm 0.14$ is consistent with the value derived by Matsunaga et al. (2006). Fixing the slope to their more accurate value results in a zero point, and implies a distance modulus to the GC of 14.51 $\\pm$ 0.12, with an additional systematic uncertainty of 0.07 mag. Similarly, from the RR Lyrae $K$-band period-luminosity relation we derive a value of 14.48 $\\pm$ 0.17 (random) $\\pm$ 0.07 (syst.). The two independent determinations are averaged to find 14.50 $\\pm$ 0.10 (random) $\\pm$ 0.07 (syst.), or 7.94 $\\pm$ 0.37 $\\pm$ 0.26 kpc. The absolute magnitude scale of the adopted period-luminosity relations is tied to an LMC distance modulus of 18.50 $\\pm$ 0.07. } {} ", "introduction": "The distance to astronomical objects is a crucial parameter, yet often very difficult to obtain with high precision. The distance to the Galactic Centre (GC) is of special importance, e.g for dynamics (Oort constants, determining distances using a rotation model), or for calibrating standard candles. The classically accepted value comes from the review by Reid (1993) and is $R_0$ = 8.0 $\\pm$ 0.5 kpc. Over the last few years the distance to the GC based on the orbit of the star called S2 around the central black-hole (BH) has caught attention. Initially, Eisenhauer et al. (2003) derived a value of 7.94 $\\pm$ 0.42 kpc which was revised by Eisenhauer et al. (2005) to 7.62 $\\pm$ 0.32 kpc having more epochs of data available. The neglect of post-Newtonian physics in these analysis may have lead to an underestimate of the distance by about $0.11 \\pm 0.02$ kpc (Zucker et al. 2006), leading to a current best estimate of 7.73 $\\pm$ 0.32 kpc (corresponding to a distance modulus (DM) of 14.44 $\\pm$ 0.09) to the GC based on the BH. On the other hand, most other recent distance determinations give a longer distance, more in line with the classical value: (1) High-amplitude delta-scuti stars give 7.9 $\\pm$ 0.3 kpc (McNamara et al. 2000); (2) RR Lyrae stars suggest a value of 8.8 $\\pm$ 0.3 kpc (Collinge et al. 2006), 8.3 $\\pm$ 1.0 kpc (Carney et al. 1995) or 8.0 $\\pm$ 0.65 kpc (Fernley et al. 1987). (3) Earlier work on the Red Clump gave a longer distance of 8.4 $\\pm$ 0.4 kpc (Paczy\\'nski \\& Stanek 1998), although Nishiyama et al. (2006) derive 7.52 $\\pm$ 0.10 (stat) $\\pm$ 0.35 (syst) kpc, and Babusiaux \\& Gilmore (2005) 7.7 $\\pm$ 0.15 kpc; (4) From a comparison of Miras found in the OGLE database in the direction of the Galactic Bulge (GB) to those in the Magellanic Clouds, Groenewegen \\& Blommaert (2005) find a distance in the range 8.5 to 9.0 kpc, in agreement with earlier work on Miras (Catchpole et al. 1999); (5) Analysis of the Hipparcos proper motions of 220 Cepheids lead to $R_0$ = 8.5 $\\pm$ 0.5 kpc (Feast \\& Whitelock 1997); (6) Modelling the observed colour-magnitude diagram in $V,I$ and $J,K$ using a population synthesis code, Vanhollebeke et al. (2008) derive a distance of 8.60 $\\pm$ 0.16 kpc. With the exception of some Red Clump based distances, the results obtained by Eisenhauer et al. imply a much shorter distance to the GC than found by most other methods, and this calls for an independent investigation of this matter. In this paper the distance to the GC is determined using \\tc\\ (hereafter P2C) discovered in the OGLE micro-lensing survey, and for which the mean $K$-band magnitude will be determined by infrared monitoring. Comparing to the calibrated P2C period-luminosity (PL) relation in the $K$-band from Matsunaga et al. (2006; hereafter M06) then provides the distance, after correction for reddening. In addition, the mean $K$-band magnitude will be determined for RR Lyrae stars that are in the field, and compared to the calibrated $K$-band PL-relation from Sollima et al. (2006). The Matsunaga et al. and Sollima et al. relations both imply an LMC DM of 18.50 as detailed in Sect.~5. In Sect.~2 the sample is discussed, and the observations are presented in Sect.~3 for the P2C and Sect.~3 for the RR Lyrae. The results are discussed in Sect.~5. ", "conclusions": "The $PL$-relation in the $K$-band of P2C in the GB is derived. The slope is found to be in agreement with that derived by M06. Fixing the slope to their more accurate value implies a DM to the GC of 14.51 with a formal error bar of 0.03. There is also the systematic error bar to consider. M06 presented $JHK$ period-luminosity relations based on 46 P2C with periods between 1.2 and 80 days in 26 Galactic globular clusters (GCs). For the absolute magnitude scale they adopted a relation between absolute $V$-magnitude of the Horizontal Branch and metallicity (Gratton et al. 2003), which in turn is calibrated using main-sequence fitting to three GCs. This calibration implies an RR Lyrae based LMC DM of 18.50 $\\pm$ 0.09 (Gratton et al. 2003). M06 show that the DM based on P2C in the LMC and their $K$-band $PL$-relation is also compatible with 18.5. They also show that there is no significant trend with metallicity over the range $-2.2 \\la$ [Fe/H] $\\la -0.5$, in agreement with theoretical predictions (Bono et al. 1997, 2003, Di Criscienzo et al. 2007), and indicating that this $K$-band $PL$-relation should be applicable for GC P2C as well. The metallicity of the P2C in the Bulge is unknown but that of RR Lyrae is estimated to be on average [Fe/H]= $-1.0$ (Walker \\& Terndrup 1991). Any difference between that metallicity and the mean metallicity of about [Fe/H]= $-1.5$ of the GCs in M06 would result in an uncertainty in the ZP of \\less 0.03 mag. Di Criscienzo et al. (2007) show that adopting a different $M_{\\rm V}$-[Fe/H] relation has a negligible effect on the derived slope of the NIR $PL$-relations. The ZP in the calibrating relation by Gratton et al. has a formal error of 0.07 and this has to be considered as a source of systematic uncertainty in the derived distance. There is other source of (random) error to consider, namely how representative this particular set of 39 stars (minus the one outlier) that defines the $PL$-relation is in view of the fact that they scatter along the line-of-sight due to the intrinsic depth of the Bulge. To simulate this, additional Monte-Carlo simulations were carried out. Random samples of 38 stars were selected from the original sample, and the $PL$-relation re-derived. The dispersion in the ZP is about 0.11 mag. This is likely a slight overestimate as in this approach the randomly drawn samples do not necessarily have the large spread in period that the true sample was selected to have. The DM to the GC we derive based on the P2C is 14.51 $\\pm$ 0.12 (random) $\\pm$ 0.07 (syst). The random error could be improved further by observing additional systems when the full OGLE-{\\sc iii} database becomes available. Based on the serendipitously observed RR Lyrae stars in the field a DM of 14.52 $\\pm$ 0.18 is derived. The error bar is for 50\\% due to the uncertainty in the adopted absolute magnitude of RR Lyra itself (Sollima et al.). Their PL-relation led to an LMC distance of 18.54 $\\pm$ 0.15. If instead we would {\\em assume} the LMC distance to be 18.50 (to be consistent with the P2C calibration) then we would find a DM of 14.48 $\\pm$ 0.13 (random) $\\pm$ 0.07 (syst), were the systematic error comes from the uncertainty in the ZP of the observed LMC PL-relation. Like for the P2C sample, there is an additional 0.11 mag systematic uncertainty due to the limited sample size. The final DM to the GC based on the $K$-band RR Lyrae stars is 14.48 $\\pm$ 0.17 (random) $\\pm$ 0.07 (syst). As the two distance estimates are derived independently, they can be averaged and the best empirical estimate of the DM to the GC based on the current data is 14.50 $\\pm$ 0.10 (random) $\\pm$ 0.07 (syst). The theoretical WIK relation gives a formal result of 14.44 (or 14.33 with anomalous reddening) with an internal error bar of 0.04 mag. A random error of 0.11 has to be added to this, due to the limited sample, and as the theoretical relation is tied to the observed relations of M06 a similar systematic error bar of 0.07 has to be considered. Although within the error bar of the purely empirical results, it brings up the question of reddening and the reddening law. An additional absorption in $K$ of 0.1 mag would bring all three methods in very good agreement. On the other hand, the reddening estimates listed in Table~1 are in excellent agreement and are based on $(J-K)$ colours (Marschall et al. 2006), $(V-I)$ (Sumi 2004), and $(V-R)$ (Popowski et al. 2003). If the absorption in $K$ were underestimated, it would also imply an underestimate of the reddening in the other maps, or a significantly higher selective reddening $A_{\\rm K}/A_{\\rm V} \\sim 0.16$ instead of 0.12. A final remark is that independent distances to some of these P2C may be obtained using surface-brightness relations (e.g. Groenewegen 2004) and the Baade-Wesselink technique. This would require better sampled $K$-band light curves than were needed for the present study and well-sampled radial velocity curves. Although observationally expensive it would give an improved understanding on the systematic error in the present analysis." }, "0801/0801.2008_arXiv.txt": { "abstract": "The generalized Chaplygin gas, which interpolates between a high density relativistic era and a non-relativistic matter phase, is a popular dark energy candidate. We consider a generalization of the Chaplygin gas model, by assuming the presence of a bulk viscous type dissipative term in the effective thermodynamic pressure of the gas. The dissipative effects are described by using the truncated Israel-Stewart model, with the bulk viscosity coefficient and the relaxation time functions of the energy density only. The corresponding cosmological dynamics of the bulk viscous Chaplygin gas dominated universe is considered in detail for a flat homogeneous isotropic Friedmann-Robertson-Walker geometry. For different values of the model parameters we consider the evolution of the cosmological parameters (scale factor, energy density, Hubble function, deceleration parameter and luminosity distance, respectively), by using both analytical and numerical methods. In the large time limit the model describes an accelerating universe, with the effective negative pressure induced by the Chaplygin gas and the bulk viscous pressure driving the acceleration. The theoretical predictions of the luminosity distance of our model are compared with the observations of the type Ia supernovae. The model fits well the recent supernova data. From the fitting we determine both the equation of state of the Chaplygin gas, and the parameters characterizing the bulk viscosity. The evolution of the scalar field associated to the viscous Chaplygin fluid is also considered, and the corresponding potential is obtained. Hence the viscous Chaplygin gas model offers an effective dynamical possibility for replacing the cosmological constant, and to explain the recent acceleration of the universe. ", "introduction": "The observations of high redshift supernovae \\cite{Pe99} and the Boomerang/Maxima/WMAP data \\cite{Ber00}, showing that the location of the first acoustic peak in the power spectrum of the microwave background radiation is consistent with the inflationary prediction $\\Omega =1$, have provided compelling evidence for a net equation of state of the cosmic fluid lying in the range $-1\\leq w=p/\\rho <-1/3$. To explain these observations, two dark components are invoked: the pressureless cold dark matter (CDM) and the dark energy (DE) with negative pressure. CDM contributes $\\Omega _{m}\\sim 0.3$, and is mainly motivated by the theoretical interpretation of the galactic rotation curves and large scale structure formation. DE is assumed to provide $\\Omega _{DE}\\sim 0.7$ and is responsible for the acceleration of the distant type Ia supernovae. There are a huge number of candidates for DE in the literature (for recent reviews see \\cite{PeRa03} and \\cite{Pa03}). One possibility are cosmologies based on a mixture of cold dark matter and quintessence, a slowly-varying, spatially inhomogeneous component \\cite{8}. An example of implementation of the idea of quintessence is the suggestion that it is the energy associated with a scalar field $Q$ with self-interaction potential $V(Q)$. If the potential energy density is greater than the kinetic one, then the pressure $p=\\dot{Q}^{2}/2-V(Q)$ associated to the $Q$-field is negative. Quintessential cosmological models have been intensively investigated in the physical literature \\cite{quint}. A different line of thought has been followed in \\cite {15,16,17}, where the conditions under which the dynamics of a self-interacting Brans$% -$Dicke (BD) field can account for the accelerated expansion of the Universe have been analyzed. Accelerated expanding solutions can be obtained with a quadratic self-coupling of the BD field and a negative coupling constant $\\omega $ \\cite{15}. Dissipative effects, including both bulk and shear viscosity, are supposed to play a very important role in the early evolution of the Universe. A cosmic fluid (pressureless and with pressure) obeying a perfect fluid type equation of state cannot support the acceleration \\cite{17}. A solution to this problem, and thus avoiding the necessity of a potential for the BD field, is to assume that some dissipative effects of bulk viscous type take place at the cosmological scale \\cite{16}. A combination of a cosmic fluid with bulk dissipative pressure and quintessence matter can drive an accelerated expansion phase of the Universe and also solve the coincidence problem (the observational fact that the energy density of cold dark matter and of $Q$-matter should be comparable today) \\cite{11}. The dynamics of a causal bulk viscous cosmological fluid filled flat homogeneous Universe in the framework of the BD theory was considered in \\cite{MaHa03}. The bulk viscous pressure term in the matter energy-momentum tensor leads to a non-decelerating evolution of the Universe. Neither CDM nor DE have direct laboratory observational or experimental evidence for their existence. Therefore it would be important if a unified dark matter - dark energy scenario could be found, in which these two components are different manifestations of a single fluid \\cite{Paetal}. A candidate for such an unification is the so-called generalized Chaplygin gas, which is an exotic fluid with the equation of state $p=-B/\\rho ^{n}$, where $B$ and $n$ are two parameters to be determined. It was initially suggested in \\cite{Ka01} with $n=1$, and then generalized in \\cite{Be02} for the case $n\\neq 1$. The Chaplygin gas also appears in the stabilization of branes in Schwarzschild-AdS black hole bulks as a critical theory at the horizon \\cite% {Ka00} and in the stringy analysis of black holes in three dimensions \\cite% {Ka98}. The Chaplygin equation of state can be derived from Born-Infeld type Lagrangians \\cite{Be02}, \\cite{No05}. This simple and elegant model smoothly interpolates between a non-relativistic matter phase ($p=0$) and a negative-pressure dark energy dominated phase. The cosmological implications of the Chaplygin gas model have been intensively investigated in the recent literature \\cite{Chco}. The Chaplygin gas cosmological model has been constrained by using different cosmological observations, like type Ia supernovae \\cite{Fa02}, the CMB anisotropy measurements \\cite{Be03}, gravitational lensing surveys \\cite{De03}, the age measurement of high redshift objects \\cite{Al03} and the X-ray gas mass fraction of clusters \\cite{Cu04}. The obtained results are somewhat controversial, with some of them claiming good agreement between the data and the Chaplygin gas model, while the rest ruling it as a feasible candidate for dark matter. In particular, the standard Chaplygin gas model with $n=1$ is ruled out by the data at a 99\\% level \\cite{Cu04}. The exact solutions of the gravitational field equations in the generalized Randall-Sundrum model for an anisotropic brane with Bianchi type I geometry, with a generalized Chaplygin gas as matter source were obtained in \\cite% {MaHa05}. The possibility of constraining Chaplygin dark energy models with current Integrated Sachs Wolfe (ISW) effect data was investigated in \\cite{GiMe06}. In the case of a flat universe the generalized Chaplygin gas models must have an energy density such that $\\Omega _{c}>0.55$ and an equation of state $w<-0.6$ at 95\\% confidence level. The extent to which the knowledge of spatial topology may place constraints on the parameters of the generalized Chaplygin gas (GCG) model for unification of dark energy and dark matter was studied in \\cite{Ber06}. By using both the Poincar\\'{e} dodecahedral and binary octahedral spaces as the observable spatial topologies, the current type Ia supernovae (SNe Ia) constraints on the GCG model parameters were examined. An action formulation for the GCG model was developed in \\cite% {BaGhKu07}, and the most general form for the nonrelativistic GCG action consistent with the equation of state has been derived. The thermodynamical properties of dark energy have been investigated in \\cite{GoWaWa07}. For dark energy with constant equation of state $w>-1$ and the generalized Chaplygin gas, the entropy is positive and satisfies the entropy bound. Observational constraints on the generalized Chaplygin gas (GCG) model for dark energy from the 9 Hubble parameter data points, the 115 SNLS Sne Ia data and the size of baryonic acoustic oscillation peak at redshift, $z=0.35$ were examined in \\cite{WuYu07}. At a 95.4\\% confidence level, a combination of the three data sets gives $% 0.67\\leq B/\\rho_0^{1+n}\\leq 0.83$ (where $\\rho_0$ is the present day energy density) and $-0.21\\leq n\\leq 0.42$, which is within the allowed parameters ranges of the GCG as a candidate of the unified dark matter and dark energy. However, the standard Chaplygin gas model ($n=1$) is also ruled out by these data at the 99.7\\% confidence level. A geometrical explanation for the generalized Chaplygin gas within the context of brane world theories, where matter fields are confined to the brane by means of the action of a confining potential, was considered in \\cite{HeSe07}. The evolution of the Universe contains a sequence of important dissipative processes, including GUT (Grand Unified theory) phase transition, taking place at $t\\approx 10^{-34}$ s and a temperature of about $T\\approx 10^{27}$ K, when gauge bosons acquire mass, reheating of the Universe at the end of inflation ($t\\approx 10^{-32}$ s), when the scalar field decays into particles, decoupling of neutrinos from the cosmic plasma ($t\\approx 1$ s, $% T\\approx 10^{10}$ K), when the temperature falls below the threshold for interactions that keep the neutrinos in thermal contact, nucleosynthesis, decoupling of photons from matter during the recombination era ($t\\approx 10$ s, $T\\approx 10^{3}$ K), when electrons combine with protons and no longer scatter the photons etc. \\cite{Ma}. The first attempts at creating a theory of relativistic dissipative fluids were those of Eckart \\cite{Ec40} and Landau and Lifshitz \\cite{LaLi87}. These theories are now known to be pathological in several respects. Regardless of the choice of the equation of state, all equilibrium states in these theories are unstable and in addition signals may be propagated through the fluid at velocities exceeding the speed of light. These problems arise due to the first order nature of the theory, that is, it considers only first-order deviations from the equilibrium leading to parabolic differential equations, hence to infinite speeds of propagation for heat flow and viscosity, in contradiction with the principle of causality. Conventional theory is thus applicable only to phenomena which are quasi-stationary, i.e. slowly varying on space and time scales characterized by mean free path and mean collision time. A relativistic second-order theory was found by Israel \\cite{Is76} and developed in \\cite{IsSt76} and \\cite% {HiLi89,HiSa91} into what is called ``transient'' or ``extended'' irreversible thermodynamics. In this model deviations from equilibrium (bulk stress, heat flow and shear stress) are treated as independent dynamical variables, leading to a total of 14 dynamical fluid variables to be determined. For general reviews on causal thermodynamics and its role in relativity see \\cite{Ma} and \\cite{Ma95}. Causal bulk viscous thermodynamics has been extensively used for describing the dynamics and evolution of the early Universe, or in an astrophysical context \\cite% {ChJa97}. It is the purpose of this paper to consider the effects of a possible existence of a bulk viscosity of the generalized Chaplygin gas on the cosmological dynamics of the Universe. The viscous effects are described by using the truncated Israel-Stewart theory \\cite{IsSt76}. By using the Laplace transformation and the convolution theorem, the second order differential equation describing the evolution of the Hubble parameter $H$ is transformed into an integral equation. The field equations are solved by means of an iterative scheme. Then the general solutions of the equations are obtained in a parametric form in the zero, first, second and $m$th order approximation, and the relevant cosmological parameters (scale factor, energy density, Hubble parameter, deceleration parameter etc.) are obtained. The scalar field interpretation of the Chaplygin gas is generalized to take into account the viscosity and dissipative effects. In order to compare the predictions of the model with the observational data we have fitted the luminosity distance-redshift relation with the latest observational data of the type Ia supernovae. The model fits well these data. From the fitting we determine both the equation of state of the Chaplygin gas, and the parameters characterizing the bulk viscosity. Even by taking into account the effect of the bulk viscosity, the $n=1$ Chaplygin gas models are ruled out by the observations. The present paper is organized as follows. The physical model and the basic equations are presented in Section II. The evolution equation for the Hubble parameter is studied in Section III, and the behavior of the cosmological parameters is obtained. The observational data have been compared with the theoretical predictions of the model in Section IV. In Section V we discuss and conclude our results. In the present paper we use a system of units so that $8\\pi G=c=1$. ", "conclusions": "In the present paper we have considered the dynamics of a bulk viscous Chaplygin gas filled flat homogeneous and isotropic universe. We have derived and formulated the evolution equations of the system, we have considered their behavior by using both analytical and numerical techniques, and we have compared the predictions of our model with the supernova data. The most attractive feature of the Chaplygin gas is that it could explain the main observational properties of the Universe without appealing to an effective cosmological constant. Generally, the obtained analytical and numerical solutions of the gravitational field equations describes an accelerating universe, with the effective negative pressure induced by the Chaplygin gas and the bulk viscous pressure driving the acceleration. From the equation of state of the Chaplygin gas with $\\gamma =0$ it follows that for the critical values $p_{c}$ and $\\rho _{c}$ of the pressure and density the parameter $w_{c}=p_{c}/\\rho _{c}$ is given by $w_{c}=-B/\\rho _{c}^{n+1}-\\Pi _{c}/\\rho _{c}$. Evaluating this relation at the present time when $\\rho _{c}=\\rho _{c0}$ gives $B=-w_{c0}\\rho _{c0}^{n+1}-\\Pi \\left( \\rho _{c0}\\right) \\rho _{c0}^{n}$. The Chaplygin gas behaves like a cosmological constant for $w_{c0}=-1$, which gives the relation between the constant $B$ and the present day value of the bulk viscous pressure as \\begin{equation} B=\\rho _{c0}^{n+1}\\left[ 1-\\frac{\\Pi \\left( \\rho _{c0}\\right) }{\\rho _{c0}}% \\right] =\\left( \\frac{3H_{0}^{2}}{8\\pi G}\\right) ^{n+1}\\left[ 1-\\frac{\\Pi \\left( \\rho _{c0}\\right) }{\\rho _{c0}}\\right] , \\end{equation}% where $H_{0}=3.24\\times 10^{-18}h$ s$^{-1}$, $0.5\\leq h\\leq 1$, is the Hubble constant \\cite{PeRa03,Pa03}. Since $\\Pi \\left( \\rho _{c0}\\right) <0$, the presence of the bulk viscous effects can significantly increase the value of $B$. By comparing the model with $\\gamma =0$ to the Gold 2006 supernova data, it turns out that a good agreement with these observations can be established for a wide range of the power $\\ s\\in \\left( 0.2,~2\\right) $ which occurs in the phenomenological laws (\\ref{csi}), which characterize the bulk viscosity coefficient. The other viscosity parameter $\\alpha $ can be obtained from the equation $3^{s-1}\\alpha H_{0}^{2s-1}=1$, and by choosing a value for the Hubble parameter. For $h=0.7$ ($% H_{0}=2.268\\times 10^{-18}$s$^{-1}$) we obtain $\\alpha =\\left( 6.2385\\times 10^{-11}\\text{s}^{-0.6},~2.8573\\times 10^{52}\\text{s}^{3}\\right) $ for the above-established range of the parameter $s$. As for the equation of state of the Chaplygin gas, by taking into account the definition of $\\lambda $, $\\lambda =B/3^{n}H_{0}^{2n+2}$, and for the same value of the Hubble parameter, the confrontation with supernova data selects the pairs $\\left( n,~B\\right) $ represented on Fig. \\ref{nB}. \\begin{figure}[tbp] \\includegraphics[height=5cm]{nB.eps} \\caption{The parameter values $B$ corresponding to the best fit values $% \\left( \\protect\\lambda ,~n\\right) $ represented in a logarithmic scale as a function of $n$.} \\label{nB} \\end{figure} Scalar fields are supposed to play a fundamental role in the evolution of the early universe. The Chaplygin gas model can be also described from a field theoretical point of view by introducing a scalar field $\\phi $ and a self interacting potential $U(\\phi )$, with the Lagrangian \\cite{Ka01,Be02}, \\cite{Bi02}, \\cite{Fa02}, \\cite{De04} \\begin{equation} L_{\\phi }=\\frac{1}{2}\\dot{\\phi}^{2}-U(\\phi ). \\end{equation} The energy density and the pressure associated to the scalar field $\\phi $ associated to the bulk viscous Chaplygin gas are given by \\begin{equation} \\rho _{\\phi }=\\frac{\\dot{\\phi}^{2}}{2}+U\\left( \\phi \\right) =\\rho , \\label{rho} \\end{equation} and \\begin{equation} p_{\\phi }=\\frac{\\dot{\\phi}^{2}}{2}-U\\left( \\phi \\right) =\\gamma \\rho -\\frac{B% }{\\rho ^{n}}+\\Pi , \\label{p} \\end{equation} respectively. The scalar field and the potential can be obtained from the equations \\begin{equation} \\phi \\left( t\\right) -\\phi _{0}=\\int_{t_{0}}^{t}\\sqrt{\\left( 1+\\gamma \\right) \\rho -\\frac{B}{\\rho ^{n}}+\\Pi }dt, \\end{equation} and \\begin{equation} U\\left( t\\right) =\\frac{1}{2}\\left[ \\left( 1-\\gamma \\right) \\rho +\\frac{B}{% \\rho ^{n}}-\\Pi \\right] , \\end{equation} respectively, where $\\phi _{0}$ is an arbitrary constant of integration. The dependence of the potential $U(\\phi )$ on the scalar field $\\phi $ is represented in Fig.~\\ref{FIG8}. \\begin{figure}[!ht] \\centering \\includegraphics{f8.eps} \\caption{The potential $U(\\phi )$ of the viscous Chaplygin gas associated scalar field as a function of the scalar field $\\phi $ for a dust universe ($\\gamma =0$), $n=0.1$, $s=1/4$, and for different values of $\\lambda _0$: $% \\lambda _0=0.01$ (solid curve), $\\lambda _0=0.03$ (dotted curve), $% \\lambda _0=0.05$ (dashed curve) and $\\lambda _0=0.07$ (long dashed curve). } \\label{FIG8} \\end{figure} In conclusion, we have found that the viscous Chaplygin gas model offers a real possibility for replacing the effective cosmological constant and to explain the recent acceleration of the universe." }, "0801/0801.0705_arXiv.txt": { "abstract": "Early universe equations of state including realistic interactions between constituents are built up. Under certain hypothesis, these equations are able to generate an inflationary regime prior to the nucleosynthesis period. The resulting accelerated expansion is intense enough to solve the flatness and horizon problems. In the cases of curvature parameter $\\kappa $ equal to $0$ or $+1$, the model is able to avoid the initial singularity and offers a natural explanation for why the universe is in expansion. All the results are valid only for a matter-antimatter symmetric universe. ", "introduction": "} The gravitational field, as described by General Relativity, couples to all types of energy: rest masses, kinetic terms and interaction terms. Relativistic cosmology is, in consequence, deeply concerned with such sources. The kinetic terms and the rest-masses are currently taken into account by assuming ideal cosmic fluids constituted by ultrarelativistic and/or non-relativistic matter. Solutions of this type are found in standard texts \\cite{nar,wein} and -- when multi-component fluids are considered -- in several papers, e.g., \\cite{Solutions}. Interaction terms are commonly used only in perturbative models. In fact, using the Boltzmann equation in a Friedmann-Robertson-Walker (FRW) background, the inhomogeneities both in the cosmic microwave background (CMB) and in the matter content \\cite{Dod,kolb} are studied, for comparison with the observational data \\cite% {Wmap,LargeScale}. Notwithstanding, interactions are not considered as direct sources of gravitation in this line of research \\cite{artigo 1}. Of course, some well-known proposals do consider interaction processes as direct sources of gravitation. They are usually related to the accelerated expansion regimes: present-day dynamics \\cite{SuNo} or inflation \\cite% {Liddle}. Nevertheless, these theories do not actually consider the fundamental interactions (electromagnetic, weak and strong) between particles in the source constituents. Indeed, in the standard inflationary approaches \\cite{Guth,AlbreSten,Linde} an accelerated expansion is obtained through self-interaction processes of scalar \\textit{inflaton} fields $\\phi _{i}\\left( x\\right) $. This self-interaction is chosen so as to produce just those features which are necessary to describe the early accelerated regime \\cite{Liddle}. The phenomenological explanations for the present-day acceleration include, among others, \\emph{(i)} the quintessence models \\cite% {CalDaveStein,Stein,Zlatev,FraRos,PeRa}, which roughly follow the same lines of the inflationary theory; \\emph{(ii)} models of matter and dark-energy unification via Chapligyn-like equations of state (EOS) \\cite{Orfeu} or through equations of the Van der Walls type \\cite{Capozzielo,Kremer}; \\emph{% (iii)} models of mass-varying-neutrino type, which couple neutrinos to a quintessence scalar field \\cite{Motta}; \\emph{(iv)} models introducing interactions in the energy conservation equation \\cite{Gabi,NelsonPinto}. A formal procedure has been recently proposed for including the (fundamental) interactions as direct sources of gravitation in the cosmological context \\cite{artigo 1}. Imported from equilibrium statistical mechanics, this formalism allows the construction of realistic equations of state for both relativistic \\cite{Dashen,Reichl} and non-relativistic systems \\cite{Pat,Beth}. Our objective here is to find primeval cosmic fluid EOS taking into account \\textit{physically realistic} interaction processes between the constituent particles, in addition to their kinetic and rest-mass terms. In particular, we shall examine their effect on the scale factor evolution. It will be shown that under certain hypothesis and approximations an early accelerated regime can be obtained as a consequence of these interacting processes. The idea of building realistic equations of state considering interaction between elementary particles in the primeval universe is not new. Actually, during the late 1970's and the beginning os the 80's a series of papers by Bugrii, Trushevsky and Beletsky \\cite{ucraone,ucratwo,ucrathree,ucrafour} discussed the construction of the high-energy EOS and their application to the pre-nucleosynthesis universe.\\footnote{% The authors are thankful to an unknown referee for calling their attention to these works.} Nevertheless, their treatment is different from the one developed here as we will make clear some pages ahead. The paper is organized as follows. Section \\ref{sec-Pns} reviews the general features of the early universe in its standard presentation \\cite% {nar,wein,Dod,kolb,Liddle}. Section \\ref{sec-SisInt} presents some results from equilibrium statistical theory of interacting systems. Specifically, the coefficients appearing in the perturbative fugacity expansions are expressed in terms of the scattering matrix operator $\\hat{S}$. In addition, the matrix $S_{2}$ describing the two-particle scattering is associated to the experimentally observable phase-shifts. The goal of Section \\ref% {sec-EoS_Pns} is to construct realistic EOS\\ for the pre-nucleosynthesis universe and discuss the hypothesis and approximations undertaken. In Section \\ref{sec-Conseq},\\ the more direct cosmological consequences coming from these equations are examined, including the effect of driving an accelerated expansion (inflationary era). Section \\ref{sec-Final} contains some final comments. Details of a too technical nature, as well as some data, have been relegated to appendices. ", "conclusions": "" }, "0801/0801.1190_arXiv.txt": { "abstract": "Post-starburst, or E+A galaxies, are the best candidates for galaxies in transition from being gas-rich and star-forming to gas-poor and passively-evolving as a result of galaxy-galaxy interactions. To focus on what E+A galaxies become after their young stellar populations fade away, we present the detailed morphologies of 21 E+A galaxies using high resolution {\\sl HST}/{\\sl ACS} and {\\sl WFPC2} images. Most of these galaxies lie in the field, well outside of rich clusters, and at least 11 (55\\%) have dramatic tidal features indicative of mergers. Our sample includes one binary E+A system, in which both E+As are tidally disturbed and interacting with each other. Our E+As are similar to early types in that they have large bulge-to-total light ratios (median $B/T$ = 0.59), high S\\'ersic indices, ($n \\gtrsim 4$), and high concentration indices ($C \\gtrsim 4.3$), but they have considerably larger asymmetry indices ($A \\gtrsim 0.04$) than ellipticals, presumably due to the disturbances within a few $r_e$ caused by the starburst and/or the galaxy-galaxy interaction. We conclude that E+As will be morphologically classified as early-type galaxies once these disturbances and the low surface brightness tidal features fade. The color morphologies are diverse, including six E+As with compact (0.4\\,--\\,1.4 kpc) blue cores, which might be local analogs of high-$z$ ellipticals with blue-cores. The large fraction (70\\%) of E+As with positive color gradients indicates that the young stellar populations are more concentrated than the old. These positive color gradients (i.e., bluer nuclei) could evolve into the negative gradients typical in E/S0s if the central parts of these galaxies are metal enhanced. Our E+As stand apart from the E/S0s in the edge-on projection of the Fundamental Plane (FP), implying that their stellar populations differ from those of E/S0s and that E+As have, on average, a \\ml that is 3.8 times smaller. The tilt of the E+A FP indicates that the variation among their stellar populations is closely tied to the structural parameters, i.e., E+As follow their own scaling relationships such that smaller or less massive galaxies have smaller {\\sl M/L}. We find a population of unresolved compact sources in nine E+As (45\\%), all of which have merger signatures. In the four E+As with suitable color data, the compact sources have colors and luminosities consistent with newly-formed star clusters. The bright end of the cluster LF is fainter in redder E+A's, suggesting that the young star clusters fade or are disrupted as the merger remnant ages. In summary, the morphologies, color profiles, scaling relations, and cluster populations are all consistent with E+As evolving ultimately into early-types, making the study of E+As critical to understanding the origin of the red sequence of galaxies. ", "introduction": "If some galaxies evolve from star-forming, gas-rich, disk-dominated galaxies (late-types) into quiescent, gas-poor, spheroid-dominated galaxies (early-types), we should find objects caught in the midst of this transformation. The best candidates are the so-called ``E+A'', ``K+A'', or \"post-starburst\" galaxies \\citep{Dressler83,Couch87} due to their combination of late- and early-type characteristics, including both a significant young stellar population (age $\\lesssim$ 1 Gyr) and a lack of on-going star formation. These galaxies have been spectroscopically identified by their strong Balmer absorption lines and absence of emission lines (e.g., [\\ion{O}{2}] and H$\\alpha$) in various environments and at all redshifts \\citep{Zabludoff96, Poggianti99, Goto03, Blake04, Tran03, Tran04}. While the cause of the abrupt end of their star formation is poorly understood, there is strong evidence that galaxy-galaxy tidal interactions or mergers trigger the starburst in many cases. First, most E+A galaxies reside in low-density environments, such as poor groups, that are similar to those of star-forming galaxies \\citep{Zabludoff96, Quintero04, Blake04, Balogh05, Goto05, Hogg06, Yan08}. Therefore, many E+As must arise from a process common in the field, such as galaxy-galaxy interactions, instead of a mechanism limited to denser, hotter environments, such as ram pressure stripping \\citep{Gunn&Gott72} or strangulation \\citep{Balogh00}. Second, a significant fraction of E+As have tidal features \\citep{Zabludoff96, Yang04, Blake04, Tran03, Tran04, Goto05}. Third, optical and {\\sl NIR} colors show that the spectral signatures of E+As require enhanced recent star formation, rather than simply a truncation of star formation in a normal spiral galaxies \\citep{Balogh05}. What will E+A galaxies become? In general, E+As are bulge-dominated, highly-concentrated \\citep{Quintero04, Tran04, Blake04, Goto05, Balogh05}, relatively gas-poor \\citep{Chang01, Buyle06}, and kinematically hot systems \\citep{Norton01}. Therefore, in a statistical sense, E+As are likely to become E/S0 galaxies. However, due to a lack of spatial resolution in previous studies, we do not know whether the detailed properties of individual E+As, e.g., their bulge fractions, color gradients, internal kinematics, and newly formed stellar clusters, are consistent with their presumed evolution into early type galaxies. Using {\\sl HST/WFPC2}, \\citet{Yang04} showed that their morphological features are consistent with a transition from late to early types, but their sample contained only the five bluest E+As galaxies from the Las Campanas Redshift Survey (LCRS; \\cite{Zabludoff96}) and thus was not representative. As a result, we still do not know whether the entire population of E+As will evolve into E/S0s, whether there is a distinguishable subclass of E+As that evolves into E/S0s, or whether E+As evolve into typical E/S0s. To answer these questions, we must understand how well E+As match the {\\it full} range of E/S0 properties. Most fundamentally, are the global morphologies (e.g., bulge-to-total light ratios and concentration) of the whole LCRS E+A sample consistent with those of E/S0s? Second, E/S0s in the local universe become redder toward their center; these negative color gradients originate from metallicity gradients \\citep{Peletier90}. In contrast, E+As exhibit a wide range of color morphologies \\citep{Yang04, Yamauchi05}. Can the color profiles of E+As evolve into those of the typical early types? Third, the number of globular clusters per unit luminosity is higher in early types than in late types \\citep{Harris&vandenBergh81}. If E+As are in transition from late to early types, one should find new star clusters formed during the starburst. Are there such clusters, and, if so, do their colors and numbers coincide with the expected evolution of the globular cluster systems of present-day E/S0s? Fourth, early-type galaxies lie on the Fundamental Plane (FP), an empirical scaling relation between the effective radius, the central velocity dispersion, and the mean surface brightness, with remarkably small scatter \\citep{Djorgovski87, Dressler87}. Will E+As lie on the same FP once they evolve? To determine whether E+As evolve into objects that are indistinguishable from the bulk of E/S0s, we present {\\sl HST/ACS} observations of the 15 remaining E+A galaxies from the \\citet{Zabludoff96} sample. We combine these with the previous {\\sl HST/WFPC2} observations of five blue E+As \\citep{Yang04} and of one serendipitously discovered E+A \\citep{Yang06}. The resulting high resolution imaging of the 21 confirmed E+A galaxies in the LCRS sample enables us to study the detailed color morphologies and the properties of the newly formed star cluster candidates at the sub-kpc scale. The detailed morphologies, color profiles, and cluster populations of many E+As are consistent with the galaxy-galaxy interaction scenario. We discuss how these features will evolve and then relate this evolution to the properties of E/S0s. Using existing kinematic data \\citep{Norton01} and our {\\sl HST} photometry, we also address whether or not the various scaling relations of E+As are consistent with those of E/S0s. This paper is organized as follows. We describe our E+A galaxy sample and the {\\sl HST/ACS} data reduction in \\S \\ref{sec:observation}. In \\S \\ref{sec:morphology}, we examine the general morphology of E+As, including a discussion of tidal features (\\S \\ref{sec:qualitative_morphology}), surface brightness profiles (\\S \\ref{sec:fitting}), structural parameters (\\S\\ref{sec:profile}), and concentration/asymmetry measures (\\S \\ref{sec:ca}). The color profiles, including a class of E+As with luminous blue cores \\citep{Yang06}, are presented in \\S \\ref{sec:color_profile}. We compare the scaling relations of E+As with those of E/S0s in \\S \\ref{sec:fp}. We present the properties of the newly-formed young star clusters in \\S \\ref{sec:cluster}. We summarize in \\S\\ref{sec:conclusion}. \\begin{figure*} \\epsscale{0.9} \\plotone{f1.low.ps} % \\caption{ ({\\it Left}) High-contrast $R$ band (\\R{702}) images show the low surface brightness tidal features. ({\\it Middle}) $R$ band images for the {\\sl WFPC2} sample (EA01AB -- EA05). ({\\it Right}) Residual $R$ band images subtracted from the smooth symmetric model components. We bound each image with 4 arcsec tickmarks and include a 4 kpc horizontal scalebar. Note the diverse morphologies of E+A galaxies: tidal and disturbed features, dusty galaxies, blue-cores, bars, and even compact star clusters. Because EA01A is too disturbed to be modeled by axisymmetric models, we restrict our analysis to EA01B and show the residual image only for EA01B in the top panel. \\label{fig:images_wfpc2}} \\end{figure*} \\begin{figure*} \\epsscale{0.9} \\plotone{f2a.low.ps} % \\caption{ ({\\it Left}) Same as for Fig \\ref{fig:images_wfpc2}, except these are {\\sl ACS} images of EA06--20. ({\\it Middle}) Two-color composite images from the $B$ and $R$ bands. ({\\it Right}) Same as for Fig \\ref{fig:images_wfpc2}, except for the {\\sl ACS} images. \\label{fig:images_acs} } \\end{figure*} \\begin{figure*} \\addtocounter{figure}{-1} \\epsscale{0.9} \\plotone{f2b.low.ps} \\caption{Continued.} \\end{figure*} \\begin{figure*} \\addtocounter{figure}{-1} \\epsscale{0.9} \\plotone{f2c.low.ps} \\caption{Continued.} \\end{figure*} ", "conclusions": "\\label{sec:conclusion} We study the detailed morphologies of 21 E+A galaxies using high resolution {\\sl HST}/{\\sl ACS} and {\\sl WFPC2} images to investigate into what E+A galaxies will evolve after their young stellar populations fade away in a few Gyr. Our findings are: \\\\ 1. The morphologies of E+As are extremely diverse, ranging across train-wrecks, barred galaxies, and blue-cores to relaxed disky galaxies. Most of these galaxies lie in the field, well outside of rich clusters, and at least 11 (55\\%) have tidal or other disturbed features. Our sample includes one binary E+A system, in which both E+As are tidally disturbed and interacting with each other. These results support the picture in which galaxy-galaxy tidal interactions or mergers are responsible for triggering the E+A phase in many cases. 2. E+As are bulge-dominated systems (median bulge fraction $B/T$ = 0.59) and their light is highly concentrated (S\\'ersic index $n \\gtrsim 5$). When dust is negligible (at least 67\\% of the time), E+As have high concentration indices ($C \\gtrsim 4.3$) consistent with those of spheroids, but considerably larger asymmetry indices ($A \\gtrsim 0.04$) than ellipticals due to structures within a few $r_e$ that presumably arise from the starburst and/or recent merger. Thus E+As would be morphologically classified as early-type galaxies once these disturbances relax and the low surface brightness tidal features dissipate or fade. 3. The color morphologies of E+As are as diverse as their structural morphologies. A large fraction (70\\%) have positive color gradients (bluer toward center), indicating that their young stellar populations are more concentrated than their older populations. We demonstrate that evolution can invert these gradients into the negative gradients typical of E/S0s if the inner parts of E+As have become more metal enriched than the outer parts due to the centralized star formation. 4. Six E+As (30\\%) exhibit compact (0.4--1.4 kpc) blue cores, which might be the local analogs of the high-$z$ elliptical blue-cores \\citep{Menanteau01a}. We discovered LINERs in three of these blue-core E+As \\citep{Yang06}, and the relationship between LINERs and blue-cores could be an important clue to what stops the star formation in E+A galaxies. 5. E+As stand apart from the E/S0 fundamental plane (FP) in the edge-on projection, implying that the stellar populations of E+As are different from that of E/S0s. E+As have, on average, a \\ml that is 3.8 times smaller than that of E/S0s. The tilt of the E+A FP indicates that the variation of the stellar populations among E+As is closely tied to their structural parameters, i.e., E+As follow their own scaling relationships such that smaller or less massive galaxies have smaller {\\sl M/L}. Such a trend arises naturally within a merger scenario, where low mass galaxies (the progenitors of low-mass E+As) have higher gas fractions \\citep{Young&Scoville91} and could produce relatively larger populations of young stars. 6. We find a population of unresolved compact sources in at least nine E+A galaxies (45\\%). The colors and luminosities of these young star cluster candidates are consistent with the ages inferred from the E+A spectra (0.01 -- 1 Gyr). The bright end of the cluster luminosity function fades as the host galaxy becomes redder, suggesting that the newly-formed young star clusters age in parallel to their host. This interpretation is confirmed by the color evolution of the cluster systems. We have now examined the full set of E+As from the Las Campanas Redshift Survey and so have representative results for local E+A galaxies. We have used high spatial resolution images to probe their detailed morphologies. We find that their properties are either consistent with those of E/S0s or, if left to evolve passively, will become like those of early-types. The morphologies, color profiles, scaling relations, and young star clusters suggest that E+As galaxies are caught in the act of transforming from late-type to early-type galaxies." }, "0801/0801.4583_arXiv.txt": { "abstract": "We study the long term evolution of magnetic fields generated by an initially unmagnetized collisionless relativistic $e^+e^-$ shock. Our 2D particle-in-cell numerical simulations show that downstream of such a Weibel-mediated shock, particle distributions are approximately isotropic, relativistic Maxwellians, and the magnetic turbulence is highly intermittent spatially, nonpropagating, and decaying. Using linear kinetic theory, we find a simple analytic form for these damping rates. Our theory predicts that overall magnetic energy decays like $(\\omega_p t)^{-q}$ with $q \\sim 1$, which compares favorably with simulations, but predicts overly rapid damping of short wavelength modes. Magnetic trapping of particles within the magnetic structures may be the origin of this discrepancy. We conclude that initially unmagnetized relativistic shocks in electron-positron plasmas are unable to form persistent downstream magnetic fields. These results put interesting constraints on synchrotron models for the prompt and afterglow emission from GRBs. ", "introduction": "The prompt emission and afterglows of gamma-ray bursts (GRBs) may be manifestations of ultrarelativistic shock waves. These shock waves may be mediated via the relativistic form of Weibel instability (Weibel 1959; Yoon and Davidson 1987; Medvedev and Loeb 1999; Gruzinov and Waxman 1999). The free energy from strong plasma anisotropy in the shock transition layer generates strong magnetic fields (with strengths comparable to the available free energy). However, these fields have very small spatial scales, i.e., the order of the plasma skin depth, $c/\\omega_p$, where $\\omega_p$ is the plasma frequency. These initially small-scale B-fields must survive for tens of thousands to millions of inverse plasma periods to serve as this source of the magnetization for synchrotron models of burst emission and afterglows (Gruzinov \\& Waxman 1999; Piran 2005ab; Katz, Keshet, \\& Waxman 2007). Whether or not these field can is an open question. Numerical and analytic studies (Kazimura {\\it et al.} 1998; Silva {\\it et al.} (2003); Frederiksen {\\it et al.} 2004; Medvedev {\\it et al.} 2005; Hededal {\\it et al.} 2005; Nishikawa {\\it et al.} 2003, 2005; Spitkovsky this proceedings) have elucidated the basic physics. The instability initially forms filaments of electric current and $B$ fields, which then merge to {\\it inverse cascade} magnetic energy to larger scales, but only in the {\\it foreshock} region. When the B-fields reach the magnetic trapping limit (Davidson {\\it et al.} 1972; Kato 2005; also see Milosavljevic, Nakar, \\& Spitkovsky 2006; Milosavljevic \\& Nakar 2006a), particle orbits become chaotic, disorganizing the filaments. The disorganized magnetic fluctuations scatter their supporting particles, which isotropizes and thermalizes the flow, within tens to hundreds of skin depths (Spitkovsky 2005). The magnetic energy peaks in this layer at $\\sim$10-20\\% of the bulk plasma flow energy. However, present simulations have not deeply followed the flow into the downstream region to explore the long term behavior of these B-fields. Thus, the question of the structure and long-term survival of the B-fields remains open (see for instance, Gruzinov \\& Waxman 1999; Gruzinov 2001ab; Medvedev {\\it et al.} 2005). In this proceeding, we discuss recent work (Chang, Spitkovsky, and Arons 2008; hereafter CSA08) which shows that this magnetic energy must rapidly decay in the downstream medium. We first describe the basic features of the downstream plasma from our numerical simulations. We then calculate the evolution (decay) of the downstream plasma using Vlasov linear response theory and then compare this evolution with simulations. While linear theory does reasonably well in estimating the decay rate of the total magnetic energy, it overestimates the damping rate of shorter wavelength modes. We discuss this discrepancy as a result of magnetic trapping. Finally, we summarize our results. ", "conclusions": "\\label{sec:discussion} We have studied the downstream evolution of magnetic turbulence in the context of a collisionless $e^+e^-$ shock both analytically and numerically. Our simulations show that the downstream region consist of nonpropagating magnetic clumps embedded in quasi-homogenous medium where the background particle distribution function is an isotropic Maxwellian. In such a background, we showed that magnetic energy will decay like $t^{-q}$ with $q\\sim 1$. However, linear theory overpredicts the decay rates at short wavelengths compared to simulations. Magnetic trapping may play an role in resolving this discrepancy. Rapid field decay puts severe constraints on GRB emission mechanism, but they may not be inconsistent with GRB observations (see Pe'er \\& Zhang 2006). Finally, if ion-electron collisionless shocks reach roughly equipartition with each other as suggest by recent large scale simulations (Spitkovsky 2008), they would reproduce the physics of the $e^{\\pm}$ shock and their B-fields would decay as well." }, "0801/0801.3530_arXiv.txt": { "abstract": "We use semi-analytic models implemented in the {\\it Millennium Simulation} to analyze the merging histories of dark matter haloes and of the galaxies that reside in them. We assume that supermassive black holes only exist in galaxies that have experienced at least one major merger. Only a few percent of galaxies with stellar masses less than $M_* < 10^{10} M_{\\odot}$ are predicted to have experienced a major merger and to contain a black hole. The fraction of galaxies with black holes increases very steeply at larger stellar masses. This agrees well with the observed strong mass dependence of the fraction of nearby galaxies that contain either low-luminosity (LINER-type) or higher-luminosity (Seyfert or composite-type) AGN. We then investigate when the major mergers that first create the black holes are predicted to occur. High mass galaxies are predicted to have formed their black holes at very early epochs. The majority of low mass galaxies never experience a major merger and hence do not contain a black hole, but a significant fraction of the supermassive black holes that do exist in low mass galaxies are predicted to have formed recently. ", "introduction": "\\label{sec:intro} By studying active galactic nuclei (AGN), we learn about the physical mechanisms that trigger accretion onto the central supermassive black holes of galaxies. When a black hole accretes, it increases in mass. By studying populations of AGN at low and at high redshifts, we hope to infer the history of how black holes build up their mass. It has been established that supermassive black holes most occur in galaxies with bulges \\citep{kormendy1995}, and that the mass of the black hole correlates with the luminosity and the stellar velocity dispersion of the host bulge \\citep{magorrian1998, ferrarese2000, gebhardt2000}. This indicates that the formation of galaxies and supermassive black holes are likely to be closely linked. In the local Universe, the fraction of bulge-dominated galaxies hosting AGN decreases at lower stellar masses \\citep{ho1997, kauffmann2003}. In order to form a black hole, it is necessary for gas to lose angular momentum and sink to the centre of the galaxy \\citep{haehnelt1993, volonteri2003}. The gravitational torques that operate during galaxy-galaxy mergers are known to be a very effective mechanism for concentrating gas at the centers of galaxies \\citep{mihos1996}. Models for AGN evolution have often assumed that black holes are formed and fuelled, and AGN activity is triggered during major mergers of galaxies \\citep{kauffmann2000, wyithe2003, croton2006}. At low and moderate redshifts, there is no conclusive observational evidence that mergers play a significant role in triggering AGN activity in galaxies. In the local Universe, \\citet{li2006} have shown that narrow line AGN do not have more close companions than matched samples of inactive galaxies. Even at intermediate redshifts (z $\\sim 0.4-1.3$), moderate luminosity AGN hosts do not have morphologies indicative of an ongoing merger or interaction \\citep{hasan2007}. The conclusion seems to be that although major mergers may be responsible for AGN activity in some galaxies, other fueling mechanisms are likely to be most important in the low redshift Universe. It has also been established that high mass black holes have largely stopped growing at early cosmic epochs, whereas low mass black holes are still accreting at significant rates today \\citep{heckman2004}. X-ray observations show that very high-luminosity AGN activity peaked at early cosmic epochs ($z \\sim 2$), while low-luminosity AGN activity peaks at lower redshifts \\citep{steffen2003,barger2005,hasinger2005}. It has been postulated that this so-called ``anti-hierarchical'' growth of supermassive black holes can be explained if there are two modes of accretion onto black holes that have very different efficiencies \\citep{merloni2004, mueller2007}. The early formation formation of ``new'' black holes may result in very luminous quasar-like events. To form a supermassive black hole, a more violent process such as a major merger may be required to funnel a large amount of gas into the central region of the galaxy. Subsequent accretion of gas onto already existing black holes may be an inefficient process and produce lower luminosity AGN \\citep{haehnelt1993, duschl2002}. The history of accretion after the black hole is formed may not necessarily be tightly linked to the dynamical history of the galaxy, but may be controlled by the accretion and feedback processes occurring in the vicinity of the black hole itself. In this work, we use the combination of the {\\it Millennium Simulation} and semi-analytic models of galaxy formation to study the fraction of galaxies that have undergone major mergers as a function of mass and cosmic epoch. We investigate whether this can be related to the demographics of black holes in the local Universe and to the apparent disappearance of the most luminous quasar activity in massive galaxies % at late times. In Sec.~\\ref{sec:simulation}, we briefly introduce the simulation we use and explain how galaxy mergers are tracked in the simulation. In Sec.~\\ref{sec:merger}, we show that if we assume that black holes only form when galaxies undergo major merging events, then most present-day low mass galaxies are predicted not not to contain black holes and hence will not host AGN. In Sec.~\\ref{sec:firstmerger}, we use the simulations to predict when galaxies of different masses have underdone their first major merger. Conclusions and discussions are presented in the final section. ", "conclusions": "\\label{sec:discussions} We analyze the merger histories of dark matter haloes and galaxies in the {\\em Millennium Simulation} and use our results to try to understand the demographics of black holes in nearby galaxies. Black holes are assumed to form only if a major merger occurs. Although a significant fraction of low mass ($< 10^{10} M_{\\odot}$) galaxies have experienced minor mergers, less than a few percent are predicted to have experienced a major merger. If our assumption that a major merger is required in order to form a black hole is correct, the majority of low mass galaxies are predicted not to contain black holes at the present day. This is one possible explanation of the observed lack of AGN in low mass galaxies \\citep{ho1997, kauffmann2003}. We also investigate when galaxies of different stellar masses are predicted to have formed their first black holes. High mass galaxies form their first black holes at very early epochs. The distribution of formation times is almost flat as a function of lookback time for low mass galaxies. This means that if a low mass galaxy has a black hole, there is a significant probability that it formed in the last few Gigyears. We also compute the number density of newly formed black holes as a function of redshift. We find that the peak number density occurs at $z \\sim2-3$, in good agreement with the observed peak in the quasar space density. More detailed predictions for how AGN of different luminosities are expected to evolve requires a more detailed physical model for how the black holes accrete gas over the history of the Universe. In addition, in certain wavebands AGN activity might be obscured by gas and dust surrounding black hole \\citep{hopkins2006}. More detailed consideration of these issues will form the basis for future work." }, "0801/0801.2911_arXiv.txt": { "abstract": "We study the tidal effects of a Kerr black hole on a neutron star in black hole-neutron star binary systems using a semi-analytical approach which describes the neutron star as a deformable ellipsoid. Relativistic effects on the neutron star self-gravity are taken into account by employing a scalar potential resulting from relativistic stellar structure equations. We calculate quasi-equilibrium sequences of black hole-neutron star binaries, and the critical orbital separation at which the star is disrupted by the black hole tidal field: the latter quantity is of particular interest because when it is greater than the radius of the innermost stable circular orbit, a short gamma-ray burst scenario may develop. ", "introduction": "During the past decades, theorists have been modelling various kinds of double compact objects since (1) they are among the most promising gravitational wave sources to be detected by ground-based and space-based laser interferometers \\cite{Detectors} and (2) they have also been invoked as possible engines of short gamma-ray bursts \\cite{GRB1} (see also \\cite{GRB,LeeGRB}) in the case of black hole-neutron star (BH-NS) and neutron star-neutron star (NS-NS) mergers. The remnants of both kinds of mergers, in fact, may result in a black hole with negligible baryon contamination along its polar symmetry axis and surrounded by a hot massive accretion disk: before the disk gas is accreted to the black hole, intense neutrino fluxes are emitted which, through energy transfer, trigger a high-entropy gas outflow off the surface of the accretion disk (``neutrino wind''); at the same time, energy deposition by $\\nu\\bar\\nu$ annihilation in the baryon-free funnel around the rotation axis, powers relativistically expanding $e^\\pm\\gamma$ jets which can give rise to gamma-ray bursts \\cite{GRB1}. The fate of BH-NS binaries, in particular, depends on the relative values of $r_{ISCO}$, the radius of the innermost stable circular orbit, and $r_{tide}$, the orbital separation at which the tidal disruption of the star by the BH occurs: if $r_{ISCO}$20~pc in front of or behind the center of Orion~OBIb. % It would be preferable to have a direct measure of the cluster's distance that is more precise than the Hipparcos distance to $\\sigma$~Ori. In a brief abstract, \\citet{garrison67} said that main~sequence fitting to 15 B stars near $\\sigma$~Ori yielded a narrow main~sequence at a distance modulus of 8.2 (440~pc). Garrison did not correct for the small values of reddening that some of the likely cluster members have. Garrison does not appear to have ever published a more detailed description of this result. In this paper we re--examine the main~sequence fitting distance for the $\\sigma$~Ori cluster using published spectroscopy and photometry for the stars that lie within 30$^{\\prime}$ of $\\sigma$~Ori~AB and have spectral types earlier than F0. ", "conclusions": "From Figure~\\ref{true_cmd} it is clear that the $\\sigma$~Ori cluster must be more distant than the nominal 350~pc Hipparcos distance for $\\sigma$~Ori. We estimate a distance of 420$\\pm$30~pc for the cluster, assuming an [Fe/H] % of $-$0.16, or 444~pc assuming solar metallicity. This is consistent with, but significantly more precise (7\\%) than the Hipparcos distance (30\\%). Our distance estimate is consistent with the estimated distance to Orion OB1b. Most of the older age estimates for the cluster assumed a distance of 350~pc. Our more tightly constrained distance shows that the cluster age must be closer to the young end of the range of the estimated ages. This places the age of the cluster in the range of 2--3~Myrs. Sixteen of the 19 stars in our sample are probable members of the $\\sigma$~Ori cluster. HD~37564 is too bright to be a cluster member. HD~294279 is too faint to be a cluster member. If HD~37333 is in fact an equal mass binary, it is likely to be a cluster member. % The existence of a tight main sequence among the O, B, and A stars of the $\\sigma$~Ori cluster suggests that any other clusters within the Orion OB1a and Orion OB1b groups (such as the 25~Ori cluster \\citep{briceno05}) should also have tight main sequences." }, "0801/0801.4165.txt": { "abstract": "We report observations of the nova RS Ophiuchi (RS Oph) using the Keck Interferometer Nuller (KIN), approximately 3.8 days following the most recent outburst that occurred on 2006 February 12. These observations represent the first scientific results from the KIN, which operates in N-band from 8 to 12.5 $\\mu$m in a nulling mode. The nulling technique is the sparse aperture equivalent of the conventional coronagraphic technique used in filled aperture telescopes. In this mode the stellar light itself is suppressed by a destructive fringe, effectively enhancing the contrast of the circumstellar material located near the star. By fitting the unique KIN data, we have obtained an angular size of the mid-infrared continuum of 6.2, 4.0, or 5.4 mas for a disk profile, gaussian profile (FWHM), and shell profile respectively. The data show evidence of enhanced neutral atomic hydrogen emission and atomic metals including silicon located in the inner spatial regime near the white dwarf (WD) relative to the outer regime. There are also nebular emission lines and evidence of hot silicate dust in the outer spatial region, centered at $\\sim$ 17 AU from the WD, that are not found in the inner regime. Our evidence suggests that these features have been excited by the nova flash in the outer spatial regime before the blast wave reached these regions. These identifications support a model in which the dust appears to be present between outbursts and is not created during the outburst event. We further discuss the present results in terms of a unifying model of the system that includes an increase in density in the plane of the orbit of the two stars created by a spiral shock wave caused by the motion of the stars through the cool wind of the red giant star. These data show the power and potential of the nulling technique which has been developed for the detection of Earth-like planets around nearby stars for the Terrestrial Planet Finder Mission and Darwin missions. ", "introduction": "Classical novae (CN) are categorized as cataclysmic variable stars that have had only one \\textit{observed} outburst - an occurrence typified by a Johnson V-band brightening of between six and nineteen magnitudes \\citep{war95}. These eruptions are well-modeled as thermonuclear runaways (TNR) of hydrogen-rich material on the surface of white dwarf (WD) primary stars that, importantly, remain intact after the event. Current theory tells us that CN can be modeled as binary systems in which a lower-mass companion -- the secondary -- orbits the WD primary such that the rate of mass transfer giving rise to the observed eruption is very low. Recurrent novae (RN) are a related class of CN which have been observed to have more than one eruption. Like CN, RN events are well-represented as surface TNR on WD primary stars in a binary system, but are thought to have much higher mass transfer rates commensurate with their greater eruption frequency. There are two types of systems that produce recurrent novae -- CVs, in which the WD accretes from a main sequence star that orbits the WD on a time scale of hours, and symbiotic stars, in which the WD accretes from a red-giant companion that orbits that WD on a time scale of years. %Dwarf novae (DN) are distinct from either of these broad classes in that their eruptions, while high in %frequency, are well-modeled as a sudden release of potential energy from a large clump of material as %it is transfered from secondary to the primary. CN and RN produce a few specific elemental isotopes by the entrainment of metal-enriched surface layers of the WD primary during unbound TNR outer-shell fusion reactions. In contrast, type Ia supernovae produce most of the elements heavier than helium in the Universe through fusion reactions leading to the complete destruction of their WD primary. Some theoretical models indicate that RN could be a type of progenitor system for supernovae. Importantly, these theories are predicated on two critical factors: 1.) the system primary must be a compact carbon-oxygen core supported solely by electron degeneracy pressure and 2.) there must be some mechanism to allow the WD mass to increase secularly towards the Chandrasekhar limit. %It does appear that the RS Ophiuchi (RS Oph) system is a member of this subclass of RN and we are %very fortunate to have it as a nearby laboratory to advance our understanding of the astrophysics of %these great engines of nucleosynthesis. The nova RS Oph has undergone six recorded episodic outbursts of irregular interval in 1898 \\citep{fle04}, 1933 \\citep{ada33}, 1958 \\citep{wal58}, 1967 \\citep{bar69}, 1985 \\citep{mor85} and now 2006. There are also two possible outbursts in 1907 \\cite[]{sch04} and 1945 \\cite[]{opp93}. All outbursts have shown very similar light curves. This system is a single-line binary, symbiotic with a red giant secondary characterized as K$5.7 \\pm 0.4$ I-II \\citep{ken87} to a K7 III \\citep{mur99} in quiescence and a white dwarf primary in a $455.72\\pm0.83$ day orbit about their common center of mass as measured using single-line radial velocity techniques \\cite[]{fek00}. \\citeauthor{opp93} examined all the outbursts and found that V band luminosity of RS Oph decreased averaged 0.09 magnitudes/day for the first 43 days after outburst. A 2-magnitude drop would then require on average 22 days, establishing RS Oph as a \u00d2fast\u00d3 nova based on the classification system of \\cite{pay57}. %:speed of rsoph % see page 130 in notebook 3 for speed calculation using AAVSO data. The most recent outburst of the nova RS Oph was discovered at an estimated V-band magnitude of 4.5 by H. Narumi of Ehime, Japan on 2006 February 12.829 UT \\cite[]{nar06}. This is 0.4 mag brighter than its historical average AAVSO V-band {\\it peak} magnitude so it is reasonable to take Feb 12.829 (JD 2453779.329) as day zero. The speed of an outburst is characterized by its $t_2$ and $t_3$ times which are the intervals in days from the visible maximum until the system has dimmed by 2 and 3 magnitudes, respectively. For this outburst the $t_2$ and $t_3$ times are 4.8 and 10.2 days, respectively. %As the WD approaches the Chandrasekhar limit the amount of mass initially ejected by the %thermonuclear runaway (TNR) decreases. It would not, therefore be unreasonable to expect the %initial V-band speed of the nova to gradually increase. Indeed it must be the case that the mass of the %WD has grown over time. If earlier in its secular evolution it was a 1.0 $M_{\\sun}$ WD and still had a %20 year recurrence time, its accretion rate would have had to be unacceptably high. %:distance to rsoph The distance to the RS Oph system is of importance to the interferometry community as it effects interpretation of astrometric data (cf. \\citet{mon06}). There has been a good deal of disagreement in the literature with a surprisingly broad range, from as near as 0.4 kpc \\citep{hac01} to as far as 5.8 kpc \\citep{pot67}. %As late as 2001, \\citeauthor{hac01} had estimated the nearer distance while early analysis of X-ray %data from the Rossi X-ray Timing Explorer confirmed a distance of 1.6 kpc \\citep{sok06}. Distance %estimates using absorption line calculations in the nova's last epoch (\\cite{hje86} \\& \\cite{sni85}) and, %using envelope expansion parallaxes from new radio observations from the current epoch \\citep %{rup08} establish a distance at the mid range. \\citet{bar08} have recently undertaken a thorough review of the various techniques that have been used to derive a distance to RS Oph and obtain a distance of $1.4^{+ 0.6}_{-0.2} $ kpc. It is this value that we adopt for astrometric calculations in this paper. %The $t_3$ and $t_2$ times noted above do constrain the distance to RS Oph and this is worth noting here. If we restrict ourselves to the use of the $t_3$ time as a more settled figure, and apply the Maximum Magnitude Rate of Decline (MMRD) expression attributed to \\citet{dow00} we obtain a range of distances of $D_{RS Oph} = 2.51 \\pm 1.28$ kpc. To this broad range we apply an additional constraint. As first noted by \\citet{cas85}, the distance to RS Ophiuchi cannot exceed approximately 2.0 kpc because it's spectrum does not contain evidence of absorption by material in the Carina arm of the Galaxy. Inclusion of the Galactic upper constraint and the MMRD lower constraint calculated above yields a range of distances to the nova RS Ophiuchi of $1.61 \\pm 0.39$ kpc. %Note: see journal #3, pages 130 - 134 The structure of this paper is as follows. We report high-resolution N-band observations of RS Oph using the nulling mode of the 85 meter baseline Keck Interferometer, beginning with a discussion of the nulling mode itself in Section 2. We discuss the observations in Section 3, and the data and analysis in Section 4. In Section 5 we introduce a new physical model of the system, which unifies many of the observations into a coherent framework. The results of this paper and those of other recent observations of RS Oph are discussed in the context of this model in Section 6. Finally, Section 7 contains a summary of our major results and conclusions. % note to Bill - we don't specifically take Lane et al into account so I changed the words \"all of the observations\" to \"many...\" ", "conclusions": "We analyzed data from the recurrent nova RS Oph for the epoch at $\\sim$ 4 days post-outburst using the new KIN instrument. These data allowed us to determine the size of the emitting region around the RS Oph at wavelengths from 8-12 $\\mu$m. By fitting the unique KIN inner and outer spatial regime data, we obtained an angular size of the mid-infrared continuum of 6.2, 4.0, or 5.4 mas for a disk profile, gaussian profile (FWHM), and shell profile respectively. The data show evidence of enhanced neutral atomic hydrogen emission located in the inner spatial regime relative to the outer regime. There is also evidence of a 9.7 $\\mu$m silicate feature seen outside of this region, which is consistent with dust that had condensed prior to the outburst, and which has not yet been disturbed by the blast wave from the nova. Our analysis of the observations, including the new ones presented in this paper, are most consistent with a new physical model of RS Oph, in which spiral shock waves associated with the motion of the two stars through the cool wind from the red giant create density enhancements within the plane of their orbital motion. Further observations are needed to clarify this new picture of the RS Oph system. One issue that has not been fully resolved is whether or not the red giant star really overflows its Roche lobe. If so a hot spot would be expected where the material streaming from the red giant envelope hits the accretion disk, and UV or X-ray observations could search for this effect. Another approach would be to observe RS Oph over several orbital cycles using infrared photometry to look for variations of the light curve due to the departure of the red giant star from spherical symmetry. High resolution spectra could also help. Confirmation of the rotational velocity of the red giant star measured by Zamanov et al. (2007) would be worthwhile, and could provide another estimate of the absolute size of the red giant star, assuming that it is co-rotating with the orbit. Further KIN observations within the next few years would also be helpful as they could show evidence of the re-establishment of the spiral shock wave, and perhaps some information about the shape of the circumstellar material and dust formation. Another epoch of HST observations would also determine the deceleration of the outflow in the two directions, i.e., within the plane of the orbit of the two stars and along the poles. Theoretical studies of the motion of the blast wave in an environment with a high density region in the plane of the orbit of the two stars are also worthwhile. In particular, it would be important to understand how the blast wave is diffracted around the red giant star and how it propagates in an medium with the periodic density enhancements in the plane due to the spiral pattern. The recurrent nova RS Ophiuchi is a rich system for the study of circumstellar matter under extreme physical conditions. Continued study will provide important insights into Type Ia supernovae, of which RS Oph may be a progenitor." }, "0801/0801.3898_arXiv.txt": { "abstract": "{We analyze the \\ha spectral variability of the rapidly-rotating K1-dwarf LQ\\,Hya using high-resolution \\ha spectra recorded during April-May 2000. Chromospheric parameters were computed from the \\ha profile as a function of rotational phase. We find that all these parameters vary in phase, with a higher chromospheric electron density coinciding with the maximum \\ha emission. We find a clear rotational modulation of the \\ha emission that is better emphasized by subtracting a reference photospheric template built up with a spectrum of a non-active star of the same spectral type. A geometrical plage model applied to the \\ha variation curve allows us to derive the location of the active regions that come out to be close in longitude to the most pronounced photospheric spots found with Doppler imaging applied to the photospheric lines in the same spectra. Our analysis suggests that the \\ha features observed in LQ~Hya in 2000 are a scaled-up version of the solar plages as regards dimensions and/or flux contrast. No clear indication of chromospheric mass motions emerges. ", "introduction": "According to the solar-stellar analogy, chromospheric stellar activity can be established by the presence of emission in the core of the Balmer \\ha line. As in the Sun, \\ha emission intensification has often been observed in surface features (plages) spatially connected with the photospheric starspots \\citep[see, e.g.,][and references therein]{Cata00,Bia06,Bia07}. Thus, the time variability of the \\ha spectral features can be used to estimate the basic properties of the emitting sources, allowing the geometry of the stellar chromosphere to be mapped. In this paper we analyze the \\ha spectral variability of the young, rapidly-rotating ($P_{\\rm rot}\\approx 1.6$\\,d) single K2-dwarf \\object{LQ\\,Hydrae} (HD\\,82558 = Gl\\,355) using 15 high-resolution spectra recorded during April-May 2000. All \\ha spectra were acquired simultaneously with the mapping lines used for the year-2000 Doppler imaging study presented in our previous paper \\citep[][hereafter Paper~I]{Kova04}. LQ\\,Hya was first recognized as a chromospherically active star through \\ion{Ca}{ii} H\\&K emission \\citep{bide81,hein81}. The \\ha absorption filled in by chromospheric emission was reported first by \\citet{feke86}, and LQ\\,Hya was classified as a BY\\,Dra-type spotted star. Variable \\ha emission-peak asymmetry was investigated by \\citet[][hereafter Paper~II]{Stra93}, who attributed it to the presence of chromospheric velocity fields in the \\ha forming layer probably surrounding photospheric spots. In our spectroscopic study in Paper~I, we presented Doppler images using the \\ion{Fe}{i}-6411, \\ion{Fe}{i}-6430, and \\ion{Ca}{i}-6439 lines for both the late April data and the early May 2000 dataset. Doppler imaging was supported by simultaneous photometric measurements in Johnson-Cousins $VI$ bands. Doppler images showed spot activity uniformly at latitudes between $-20$\\degr and $+50$\\degr, sometimes with high-latitude appendages, but without a polar spot. Comparing the respective maps from two weeks apart, rapid spot evolution was detected, which was attributed to strong cross-talks between the neighboring surface features through magnetic reconnections. In Table~\\ref{chart} we give a summary of the stellar parameters as they emerged from Paper~I. \\begin{table} \\caption{Astrophysical data for LQ\\,Hya \\citep[adopted from][]{Kova04}} \\label{chart} \\centering \\begin{tabular}{ll} \\hline\\hline \\noalign{\\smallskip} Parameter & Value \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} Classification & K2 V \\\\ Distance (Hipparcos) & 18.35$\\pm$0.35 pc \\\\ $(B-V)_{\\rm Hipparcos}$ & 0.933$\\pm$ 0.021 mag \\\\ $(V-I)_{\\rm Hipparcos}$ & 1.04$\\pm$ 0.02 mag \\\\ Luminosity, $L$ & 0.270$\\pm$0.009~L$_\\odot$ \\\\ $\\log g$ & 4.0$\\pm$0.5 \\\\ $T_{\\rm eff}$ & 5070$\\pm$100 K \\\\ $v\\sin i$ & 28.0$\\pm$1.0 km/s \\\\ Inclination, $i$ & 65$^{\\circ}$$\\pm$10$^{\\circ}$ \\\\ Period, $P_{\\rm rot}$ & 1.60066$\\pm$0.00013 days \\\\ Radius, $R$ & 0.97$\\pm$0.07 $R_\\odot$ \\\\ Mass & $\\approx$ 0.8 $M_\\odot$ \\\\ Age & $\\approx$ ZAMS \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\end{table} The observations are again briefly presented in Sect.~\\ref{data} and the method used for the \\ha spectral study is described in Sect.~\\ref{sec:analysis}. The results are presented in Sect.~\\ref{results}. Since the \\ha observations of this paper were included in the spectroscopic data used in Paper~I to reconstruct the year-2000 Doppler images, in Sect.~\\ref{disc} we take the opportunity to compare the photospheric features with the contemporaneous H$\\alpha$-emitting regions. ", "conclusions": "\\label{disc} To obtain information about the surface location of the active regions in the chromosphere of LQ~Hya, we applied a simple geometric {\\it plage} model to the rotationally modulated chromospheric emission. This method had been described and successfully applied to the \\ha modulation curves of several other active stars \\citep[e.g.,][]{Fra00, Bia06, Fra07}. On LQ~Hya, two circular bright spots (plages) are normally sufficient for reproducing the observed variations within the data errors \\citep[cf.][]{Fra00,KoBa97}. In our model the flux ratio between plages and the surrounding chromosphere ($F_{\\rm plage}/F_{\\rm chrom}$) is a free parameter. Solar values of $F_{\\rm plage}/F_{\\rm chrom}\\,\\approx\\,2$, as deduced from averaging many plages in H$\\alpha$ \\citep[e.g., ][]{Elli52,LaBo86,Ayres86}, are too low to model the high amplitudes of H$\\alpha$ $\\Delta EW$ curves in very active stars \\citep{Bia06,Fra07}. In fact, extremely large plages, covering a significant fraction of the stellar surface, would be required with such a low flux ratio, and they could not reproduce the observed modulation. In order to achieve a good fit, a flux ratio of 5, which is also typical of the brightest parts of solar plages or of flare regions \\citep[e.g., ][]{Sves76,Ziri88}, was adopted. The solutions essentially provide the longitude of the plages and give only rough estimates of their latitude and size. The information about the latitude of surface features can be recovered from the analysis of spectral-line profiles broadened by the stellar rotation, as we did for the spots of LQ~Hya with Doppler imaging in Paper I. A similar technique cannot reach the same level of accuracy when applied to the \\ha profile whose broadening is dominated by chromospheric heating effects, which are particularly efficient in the active regions \\citep[e.g.,][]{Lanzafame00}. We searched for the best solution by varying the longitudes, latitudes, and radii of the active regions. The radii are strongly dependent on the assumed flux contrast $F_{\\rm plage}/F_{\\rm chrom}$. Thus, only the combined information between plage dimensions and flux contrast, i.e. some kind of plage ``luminosity'' in units of the quiet chromosphere ($L_{\\rm plage}/L_{\\rm quiet}$), can be deduced as a meaningful parameter. Note also that we cannot estimate the true quiet chromospheric contribution (network), since the H$\\alpha$ minimum value, $\\Delta EW_{\\rm quiet}=0.84$\\,\\AA, could still be affected by a homogeneous distribution of smaller plages. However, such an approximation seems valid for LQ~Hya because our aim was only to compare the spatial distribution of the main surface inhomogeneities at chromospheric and photospheric levels. We also searched for solutions with three active regions, finding a small but possibly significant increase in the goodness of the fit, with the $\\chi^2$ passing from 5.77 to 4.45 (see Fig.~\\ref{fig:rotmod}). The model with only two plages requires features at rather high latitudes (Table~\\ref{tab:param_plages}) to explain the nearly flat and long-lasting maximum, while a 3-plage model allows placing plages at lower latitudes, in better agreement with the spot locations from Doppler imaging. However, the longitudes of the two largest plages do not change very much (less than 15 degrees). The resulting plage longitudes are in good agreement with the Doppler results in Paper\\,I; i.e., photospheric minima at 0.6 and 0.9 with the largest/coolest photospheric spots correspond to the most luminous chromospheric phases, and vice versa, the brightest photometric phase at 0.3 overlap with the lowest chromospheric \\ha emission, again supporting the paradigm that photospheric spots are physically connected with chromospheric plages (Fig.~\\ref{fig:palle}). The synthetic \\ha EW curve for the model with three plages is plotted in the middle panel of Fig.~\\ref{fig:rotmod} as a continuous line superimposed on the data (dots), while the 2-plage solution is displayed by a dotted line. Both of them reproduce the observed rotational modulation quite well. With the same models we were able to calculate the radial velocity shift between the \\ha synthetic emission profile, resulting from the quiet chromosphere plus the visible plages, and the photosphere. It is clear in the bottom panel of Fig.~\\ref{fig:rotmod} that the amplitude of the theoretical $\\Delta V_{\\rm em}$ curves, both for a 2-plage and a 3-plage model, is very low, consistent with the values derived from the residual profiles ($\\Delta V_{\\rm em}^{\\rm res}$) and in total disagreement with the velocity shifts derived from the raw spectra. This strongly supports the spectral synthesis method we used to evaluate the chromospheric emission in the \\ha line. We would like to outline the rotational modulation of other features observed in the \\ha profile, such as the peak separation, which is related to the chromospheric electron density (Sect.~\\ref{results}) and the asymmetry of blue/red peak intensity (Fig.\\,\\ref{fig:Ne}). The chromosphere of LQ~Hya displays a higher electron density above active regions. The blue emission peak is stronger than the red one ($\\frac{F_{\\rm R}}{F_{\\rm V}}< 1$) in the ``quiet\" chromosphere of LQ~Hya, while nearly equal peaks ($\\frac{F_{\\rm R}}{F_{\\rm V}}\\simeq 1$) tend to be observed when the most active regions are in the visible hemisphere. Asymmetry in the peaks of an emission line with a central reversal has been frequently observed for \\ion{Ca}{ii} K line both in the Sun \\citep[e.g.,][and reference therein]{Ziri88,Ding98} and in cool stars \\citep[e.g.,][and reference therein]{Montes00}.\t% The blue asymmetry is more frequently observed in the solar chromosphere and is commonly attributed to the propagation of acoustic waves \\citep[e.g.,][]{Cram76} with an upward velocity on the order of 10 \\kms\\ in the layer in which the K$_2$ emission peaks are formed or a downward motion in the layer producing the central reversal K$_3$ \\citep[e.g.,][]{Durrant76}. \\citet{Oranje83} has shown that the average \\ion{Ca}{ii}~K emission profile in solar plages has a completely different shape from the surrounding chromosphere, with the red peak slightly stronger than the blue one. This could reflect the different physical conditions and velocity fields in active regions compared to the quiet chromosphere. A similar analysis cannot be made for the Sun as a star in H$\\alpha$, because this line is an absorption feature in the quiet chromosphere and only a filling in is observed in plages. However, similar changes of the \\ha line profile can be expected in the plages of any stars that are more active than the Sun. Thus, the rotational modulation of $\\frac{F_{\\rm R}}{F_{\\rm V}}$ is consistent with the presence of plages in the chromosphere of LQ~Hya. These results suggest scaled-up versions (regarding size and brightness) of solar-type plages for the \\ha features observed in LQ~Hya in 2000, without any clear evidence of strong mass motions in its chromosphere. \\begin{figure}[tbh] \\vspace{0.5cm} \\begin{center} \\includegraphics[width=2.5cm]{9058f4a.eps}\t% \\hspace{0.4cm} \\includegraphics[width=2.5cm]{9058f4b.eps}\t% \\hspace{0.4cm} \\includegraphics[width=2.5cm]{9058f4c.eps}\t% \\includegraphics[width=2.5cm]{9058f4d.eps}\t% \\hspace{0.4cm} \\includegraphics[width=2.5cm]{9058f4e.eps}\t% \\hspace{0.4cm} \\includegraphics[width=2.5cm]{9058f4f.eps}\t% \\includegraphics[width=2.5cm]{9058f4g.eps}\t% \\hspace{0.4cm} \\includegraphics[width=2.5cm]{9058f4h.eps}\t% \\hspace{0.4cm} \\includegraphics[width=2.5cm]{9058f4i.eps}\t% \\end{center} \\caption{Schematic representation of the surface features of LQ~Hya in the chromosphere as reconstructed by the 2-plage {\\it (top)} and by the 3-plage models {\\it (middle)} and in the photosphere {\\it (bottom)}. In these maps, the star is rotating counterclockwise. Dominant features are found at similar longitudes at both chromospheric and photospheric levels. } \\label{fig:palle} \\end{figure} \\begin{table} \\caption{Plage parameters for LQ~Hya in April-May 2000.}\t% \\label{tab:param_plages} \\begin{center} \\begin{tabular}{crcc} \\hline \\hline \\noalign{\\smallskip} Radius & Lon. & Lat. & $\\frac{F_{\\rm plage}}{F_{\\rm chrom}}$ \\\\ ($\\degr$) & ($\\degr$) & ($\\degr$) & \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\multicolumn{4}{c}{2 Plages Solution}\\\\ \\hline \\noalign{\\smallskip} 15.5 & 215 & 35 & 5 \\\\ 20.0 & 353 & 66 & 5 \\\\ \\hline \\noalign{\\smallskip} \\multicolumn{4}{c}{3 Plages Solution}\\\\ \\hline \\noalign{\\smallskip} 13.0 & 210 & 20 & 5 \\\\ 13.0 & 8 & 15 & 5 \\\\ 9.5 & 280 & 15 & 5 \\\\ \\hline \\hline \\end{tabular} \\end{center} \\end{table}" }, "0801/0801.1065_arXiv.txt": { "abstract": "Strong selection effects are present in observational samples of cataclysmic variables (CVs), complicating comparisons to theoretical predictions. The selection criteria used to define most CV samples discriminate heavily against the discovery of short-period, intrinsically faint systems. The situation can be improved by selecting CVs for the presence of emission lines. For this reason, we have constructed a homogeneous sample of CVs selected on the basis of H$\\alpha$ emission. We present discovery observations of the 14 CVs and 2 additional CV candidates found in this search. The orbital periods of 11 of the new CVs were measured; all are above 3~h. There are two eclipsing systems in the sample, and one in which we observed a quasi-periodic modulation on a $\\sim 1\\,000\\,\\mathrm{s}$ time-scale. We also detect the secondary star in the spectrum of one system, and measure its spectral type. Several of the new CVs have the spectroscopic appearance of nova-like variables (NLs), and a few display what may be SW Sex star behaviour. In a companion paper, we discuss the implications of this new sample for CV evolution. ", "introduction": "Cataclysmic variable stars (CVs) are semi-detached binary stars, consisting of a white dwarf primary accreting from an approximately main-sequence companion (see e.g. \\citealt{bible}). The mass transfer is caused by orbital angular momentum loss, which also drives the secular evolution of CVs. At long orbital periods ($P_{orb}\\ga3\\,\\mathrm{h}$), angular momentum loss through magnetic braking is thought to dominate over that resulting from gravitational radiation. According to the `disrupted magnetic braking' model of CV evolution, magnetic braking stops when $P_{orb}$ reaches $\\simeq 3\\,\\mathrm{h}$, the secondary loses contact with its Roche lobe, and gravitational radiation is left as the only angular momentum loss mechanism (e.g. \\citealt{RobinsonBarkerCochran81}; \\citealt{RappaportVerbuntJoss83}; \\citealt{SpruitRitter83}). The secondary only regains contact with its Roche lobe when gravitational radiation has decreased $P_{orb}$ to $\\simeq 2\\,\\mathrm{h}$. This model explains the pronounced drop in the number of CVs at $2~\\mathrm{h} \\la P_{orb}\\la3$~h, called the period gap. Mass loss from the secondary increases its thermal time-scale, so that, for CVs below the period gap, this eventually exceeds the mass-transfer time-scale (even though the mass loss time-scale increases as $M_2$ decreases). When this happens, the secondary is not able to decrease its radius rapidly enough in response to mass loss, so that the orbital evolution slowly moves back through longer periods (e.g. \\citealt{Paczynski81}; \\citealt{PaczynskiSienkiewicz81}; \\citealt{RappaportJossWebbink82}). CVs in this final phase of evolution, where $P_{orb}$ is increasing, are referred to as `period bouncers'. The reversal in the direction of change in $P_{orb}$ causes the period minimum---a sharp cut-off in the $P_{orb}$ distribution of hydrogen-rich CVs at about 76~min. Magnetic braking is much more efficient than gravitational radiation at removing angular momentum---the resulting mass transfer rates ($\\dot{M}$) above the gap are roughly 10 to 100 times larger than below the gap (e.g. \\citealt{Patterson84}). This implies that long-period CVs are intrinsically bright, and also that the long-period phase of the evolution of a CV is very short-lived. The majority of CVs should therefore be short-period, low-$\\dot{M}$ (and therefore intrinsically faint) systems. In fact, even the time taken to evolve from the bottom of the period gap to the period minimum is short compared to the age of the Galaxy, so that most CVs should be period bouncers. Population synthesis studies predict the relative sizes of the long-period, short-period, and period bouncer populations (roughly $1:30:70$), as well as the absolute size of the Galactic CV population (e.g. \\citealt{Kolb93}; \\citealt{HowellRappaportPolitano97}). However, the predictions of relative sizes of sub-populations, as well as the overall space density of CVs, have been disputed on observational grounds (e.g. \\citealt{Patterson84}; \\citealt{Patterson98}; \\citealt{GansickeHagenEngels02}; \\citealt{PretoriusKniggeKolb07}; \\citealt{PretoriusNEP}). A meaningful quantitative comparison between observations and theory is only possible if observational selection effects can be accounted for. This requires that the criteria for inclusion in the observational sample are well-defined, and preferably homogeneous. Furthermore, the typical intrinsic brightness of the lowest-$\\dot{M}$ CV population that can be probed by a given survey clearly depends on the survey flux limit. If this limit is too bright, the survey has no sensitivity to any but the intrinsically brightest (and rarest) CVs. Several complete, uniformly selected CV samples already exist, e.g. those that resulted from the Palomar Green Survey, the \\emph{ROSAT} Bright Survey, and the \\emph{ROSAT} North Ecliptic Pole Survey (\\citealt{Ringwald93}; \\citealt{SchwopeBrunnerBuckley02}; \\citealt{PretoriusNEP}). Ongoing surveys, e.g. the Hamburg Quasar Survey, and the Sloan Digital Sky Survey (\\citealt{GansickeHagenEngels02}; \\citealt{AungwerojwitGansickeRodriguez-Gil06}; \\citealt{sdsscvs1}, 2003, 2004, 2005, 2006, 2007) are in the process of producing similarly well-defined, but much deeper, samples. All CV samples are affected by a flux-limit\\footnote{This may not be important for the discovery of classical novae, but it is for recovering them and measuring periods.}, which already implies a bias against intrinsically faint systems. But it is in fact very difficult to construct a sample that is reasonably large and deep, and purely flux-limited. Of the surveys mentioned above, only the very shallow \\emph{ROSAT} Bright Survey, and the very small \\emph{ROSAT} North Ecliptic Pole Survey do not contain a blue cut, in addition to a flux limit. The surveys that incorporate a blue selection are biased against low-$\\dot{M}$ CVs, since these systems are not only intrinsically faint, but also relatively red. The single property through which the most CVs have been discovered is large amplitude variability. It is not easy to quantify the selection effects acting on the CVs discovered in this way, but they certainly also favour intrinsically brighter systems, since low-$\\dot{M}$ CVs undergo less frequent outbursts. \\cite{Gansicke05} reviews all surveys that have discovered sizable samples of CVs. The presence of Balmer emission lines in the spectra of most CVs provides an alternative to the commonly used blue- and variability-based selection techniques. Selecting CVs for line emission discriminates only against the discovery of the intrinsically rare CVs with very bright, optically thick discs. In fact, there is a well known (and theoretically expected) anti-correlation between the equivalent widths (EWs) of Balmer emission lines and the luminosity of CVs (\\citealt{Patterson84}; \\citealt{WithamKniggeGansicke06}; we will take EWs of emission lines as positive throughout). A few surveys have already exploited the promise of discovering CVs via emission lines. CVs are selected from the Hamburg Quasar Survey based in part on H$\\beta$ emission. One of the selection criteria for CVs from the Cal\\'{a}n-Tololo Survey is Balmer emission lines \\citep{TappertAugusteijnMaza04}. The INT Photometric H$\\alpha$ Survey of the Northern Galactic Plane (IPHAS) is currently being used to find CVs (\\citealt{WithamKniggeGansicke06}; \\citealt{WithamKniggeAungwerojwit07}), and \\cite{RogelLuggerCohn06} describe another H$\\alpha$-based search for CVs. H$\\alpha$ emission has also been used to identify CV candidates in globular clusters (e.g. \\citealt{BailynRubensteinSlavin96}; \\citealt{CoolGrindlayCohn95}) The AAO/UKST SuperCOSMOS H$\\alpha$ Survey (SHS; \\citealt{Parker05}) is currently the best available southern hemisphere resource for identifying H$\\alpha$ emission line point sources. One CV has previously been discovered in this survey, partly on the basis of H$\\alpha$ emission \\citep{HowellMasonHuber02}. We have carried out spectroscopic follow-up of sources in this survey with $R<17.5$, selected for H$\\alpha$ emission. This has allowed us to construct a small, homogeneous sample of CVs. We present discovery spectra of 16 new CV candidates selected from the SHS, as well as time-resolved observations that have confirmed 14 of these systems as CVs, and yielded orbital period measurements for 11 of the new CVs. The sample will be used to derive constraints on CV evolution theory in a subsequent paper (\\citealt{HalphaII}; hereafter Paper II). ", "conclusions": "\\begin{tabular}{@{}llllllllllllll@{}} \\hline Object/coordinates & $B$ & $R$ &$R_2$ &$R_1$ & $I$ & $J$ & $H$ & $K_S$ & $R-\\mathrm{H}\\alpha$ & $\\mathrm{EW}(\\mathrm{H}\\alpha)/\\mathrm{\\AA}$ & $P_{orb}/\\mathrm{h}$& $d_l/\\mathrm{pc}$ & Notes \\\\ \\hline 073418.56-170626.5 & 17.8 & 16.7 & --- & 18.3 & 17.4 & 16.23 & 15.93 &$>$14.9& 0.47 & 61(6) & 3.18542(5)& 320 & 1,2,3\\\\ 074208.23-104932.4 & 16.5 & 15.3 & 15.4 & 16.0 & 14.3 & 14.26 & 13.59 & 13.36 & 1.21 & 63(3), 45(1) & 5.706(3) & 490 & 4,5\\\\ 074655.48-093430.5 & 13.8 & 14.0 & 13.6 & 13.4 & 13.3 & 13.47 & 13.29 & 13.19 & 0.46 & 27(1), 23.6(3)& 3.3984(4) & 180 & 6\\\\ 075648.83-124653.5 & 18.3 & 16.0 & 17.0 & 17.0 & 16.2 & 15.39 & 14.73 & 14.61 & 0.61 & 43(7), 47.1(6)& --- & --- & 7\\\\ 092134.12-593906.7 & 17.3 & 16.8 & 16.7 & 16.9 & 16.7 & 16.49 & 16.20 &$>$15.8& 0.53 & 56(6), 43.7(4)& 3.041(9) & 530 & 8\\\\ 092751.93-391052.3 & 16.4 & 16.2 & 16.8 & 16.3 & 16.1 & 15.72 & 15.54 &$>$15.7& 0.35 & 15(2), 20.0(3)& 4.1(3) & 770 & 9\\\\ 094409.36-561711.4 & 17.5 & 16.2 & --- & 18.4 & 16.7 & 15.72 & 15.40 & 14.96 & 0.99 & 66(4), 67(1) & 4.506(4) & 520 & 1,2,3,8\\\\ 102442.03-482642.5 & 17.1 & 16.4 & 17.6 & 16.4 & 16.2 & 15.83 & 15.57 & 15.41 & 0.94 &125(2), 122(1) & 3.673(6) & 580 & 3\\\\ 103135.00-462639.0 & 17.4 & 16.3 & 17.2 & 16.9 & 16.3 & 16.04 & 15.65 & 15.52 & 0.39 & 35(4), 19.4(3)& 3.76(2) & 630 & 9\\\\ 103959.96-470126.1 & 17.0 & 16.4 & 16.9 & 16.7 & 16.8 & 15.87 & 15.69 & 15.63 & 0.95 & 31(3), 20.2(5)& 3.785(5) & 670 & 9\\\\ 112921.67-535543.6 & 16.3 & 15.5 & 15.9 & 16.1 & 15.3 & 15.27 & 15.17 & 15.12 & 0.33 & 23(2), 17.4(5)& 3.6851(4) & 510 & 9\\\\ 115927.06-541556.2 & 18.6 & 17.4 & 17.6 & 17.5 & 17.2 & --- & --- & --- & 0.57 & 57(15) & --- & --- & 10\\\\ 122105.52-665048.8 & 17.8 & 17.3 & 17.5 & 16.9 & 17.4 & 16.43 &$>$16.3&$>$16.4& 0.58 & 22(3), 66.0(5)& --- & --- & 7,9,11\\\\ 130559.50-575459.9 & 17.2 & 16.5 & 16.4 & --- & 16.2 & 15.99 & 15.55 &$>$14.9& 0.59 & 41(4), 58.3(5)& 3.928(13) & 500 & 8\\\\ 163447.70-345423.1 & 16.8 & 16.7 & 16.7 & 17.0 & 16.3 & 15.88 & 15.42 &$>$15.7& 0.61 & 36(5), 34(3) & --- & --- & 3,9\\\\ 190039.83-173205.5 & 17.6 & 16.8 & 17.3 & 16.9 & 17.8 & 15.66 & 15.45 & 15.11 & 0.34 & 26(5) & --- & --- & 9,10\\\\ \\hline \\end{tabular} \\noindent Notes: (1) Eclipsing system. (2) Possibly an SW Sex star. (3) Large amplitude variability. (4) Secondary star spectral type M$0\\pm1$. (5) \\emph{ROSAT} source. (6) Bowen blend detected. (7) Highly variable H$\\alpha$ line profile. (8) Strong He\\,{\\scriptsize II}\\,$\\lambda$4686 emission. (9) Spectroscopic appearance of a nova-like variable (NL). (10) Classification as a CV not certain. (11) Probable quasi-periodic oscillation (QPO) detected in one light curve. \\hfill } \\end{table*} The SHS is a photographic survey performed with the UK Schmidt Telescope (UKST), and scanned by a digitizing machine called SuperCOSMOS. The survey imaged the southern Galactic plane in $R$ and H$\\alpha$, down to a limiting magnitude of $\\simeq 20.5$ (see \\citealt{Parker05} for a detailed description). $I$-band photometry from an older UKST survey is included with the H$\\alpha$ and $R$ data and was used in our target selection. We selected objects that are clear H$\\alpha$ excess outliers in the $R-\\mathrm{H}\\alpha$ vs $R-I$ colour-colour plane for spectroscopic follow-up. Most of the target sample was restricted to objects brighter than $R=17.0$, but we also considered objects with $17.0 \\le R < 17.5$ for one of our identification spectroscopy runs. Two CV candidates were found in the fainter magnitude bin. The selection criteria and sample completeness will be discussed in more detail in Paper II. Fig.~\\ref{fig:findingcharts} gives finding charts for the new CVs and CV candidates turned up by the search, and Table~\\ref{tab:summary} lists J2000 coordinates, broad band magnitudes, H$\\alpha$ excesses and equivalent widths, orbital periods, and lower limits on distances. We will name objects using their right ascension as `H$\\alpha$hhmmss'. Where two values of $\\mathrm{EW}(\\mathrm{H}\\alpha)$ are given, the first is obtained from the identification spectrum, and the second from the average of higher resolution spectra of that system. Errors in the equivalent widths were estimated using the method described by \\cite{HowarthPhillips86}. Near-infrared magnitudes are from the Two Micron All Sky Survey (2MASS; \\citealt{2mass})\\footnote{H$\\alpha$092751 was not detected in $K_S$ by 2MASS; the 95\\% confidence lower limit on $K_S$ is given as 16.65. This implies either very unusual near-IR colours for the source, or variability between the different IR images (which were taken on the same night). However, the number-magnitude count of sources detected in a 5 arc min radius around the position of H$\\alpha$092751 indicates that the magnitude limit of the $K_S$-band image is about 15.7. We will adopt this more conservative limit.}. The optical magnitudes listed in Table~\\ref{tab:summary} are from SuperCOSMOS scans of the UKST Blue Southern and Equatorial Surveys ($B$), the UKST Red Southern and Equatorial Surveys ($R_2$), the ESO Schmidt Telescope Red Southern Survey ($\\delta < -17.5^\\circ$) or (for $-17.5^\\circ < \\delta < +2.5^\\circ$) the Palomar-I Oschin Schmidt Telescope (POSS-I) Red Southern Extension ($R_1$), and the UKST near Infrared Southern Survey ($I$). The orbital period measurements are discussed in Section~\\ref{sec:results}, and lower limits on the distances to our sources ($d_l$) are derived in Section~\\ref{sec:distances}. \\begin{figure*} $\\begin{array}{c@{\\hspace{0.1cm}}c@{\\hspace{0.1cm}}c@{\\hspace{0.1cm}}c} \\includegraphics[width=43mm]{fig01_01.eps} & \\includegraphics[width=43mm]{fig01_02.eps} & \\includegraphics[width=43mm]{fig01_03.eps} & \\includegraphics[width=43mm]{fig01_04.eps} \\\\ \\includegraphics[width=43mm]{fig01_05.eps} & \\includegraphics[width=43mm]{fig01_06.eps} & \\includegraphics[width=43mm]{fig01_07.eps} & \\includegraphics[width=43mm]{fig01_08.eps} \\\\ \\includegraphics[width=43mm]{fig01_09.eps} & \\includegraphics[width=43mm]{fig01_10.eps} & \\includegraphics[width=43mm]{fig01_11.eps} & \\includegraphics[width=43mm]{fig01_12.eps} \\\\ \\includegraphics[width=43mm]{fig01_13.eps} & \\includegraphics[width=43mm]{fig01_14.eps} & \\includegraphics[width=43mm]{fig01_15.eps} & \\includegraphics[width=43mm]{fig01_16.eps} \\end{array}$ \\caption {$4' \\times 4'$ finding charts of the new CVs and CV candidates, made using UKST $R$-band plates digitized by SuperCOSMOS. The charts are labelled with the right ascension of the sources. North is at the top and east to the left in all images. } \\label{fig:findingcharts} \\end{figure*} We have discovered 16 CV candidates by selecting emission line objects from the SHS, and obtained additional observations for 14 of these systems, confirming their CV nature. Orbital periods were measured for 11 of the new CVs. All of these are long-period systems, and most have orbital periods in the range 3 to 4 h. The periods are listed in Table~\\ref{tab:summary}. Note that 6 of these periods were determined from aliased radial velocity curves, and may therefore only be correct to within $\\sim 10$\\% (the errors given in the table are for the strongest alias only). The aim of this study is to construct a new CV sample with uniform selection criteria. Although the bright flux limit implies that the sample is biased against intrinsically faint CVs, there is no explicit second selection cut that compounds this bias (but note that there is an effective $R-I$ cut, which will be discussed in Paper II). It is also the largest H$\\alpha$-selected CV sample constructed to date. The blue, weak-lined spectra of H$\\alpha$092751 and H$\\alpha$112921 imply that they are probably both NLs. Given the orbital periods of our CVs, it is likely that many of them are SW Sex stars. Specifically, we have reason to suspect that H$\\alpha$073418, H$\\alpha$094409, H$\\alpha$075648, and H$\\alpha$122105 are SW Sex stars, but confirmation is needed for all of them. H$\\alpha$075648 and H$\\alpha$122105 are unusual in displaying large variations in line profiles, which prevented us from measuring orbital periods. H$\\alpha$073418, H$\\alpha$094409, H$\\alpha$102442, and H$\\alpha$163447 have been observed to display large amplitude variability. The secondary star is detected in the spectrum of H$\\alpha$074208. We measure a spectral type of M$0\\pm1$, and estimate that the secondary contributes between 40 and 80\\% of the flux in the wavelength range from $\\simeq5\\,700$ to $\\simeq7\\,200\\,\\mathrm{\\AA}$." }, "0801/0801.4530_arXiv.txt": { "abstract": " ", "introduction": "Be stars are well known to be variable on virtually all timescales, reaching from minutes to dozens of years. For the study of the latter, long term data collections as homogeneous as possible are necessary. The professional astronomer, however, is often hampered in the study of intermediate- to long-term time scale processes as in Be stars. The reasons are the observational practices usually employed at professional observatories, which typically are not suited for observing a bright object with execution times of a few minutes only about every other week for several seasons; as well as the funding timescales, making it hard to start the collection of a long-term database that does not promise a significant number of publications within the first few years. On the other hand, the interpretation of time-limited observations with professional resources, such as interferometers, polarimeter, or high-resolution spectrographs, in almost all cases can profit from the knowledge of the disc state in the course of the long-term evolution. The problems in long-term data acquisition for the professional astronomer, however, open a promising field for the dedicated amateur. Amateur spectrographs at relatively small telescopes of about 20\\,cm diameter, equipped with CCD-detectors meanwhile reach resolution powers well above 10\\,000 and and are sensitive enough to reach many of the brighter Be stars. This work describes a database worth of more than five years of observation of the Be star $\\zeta$ Tau. The observational data will be made available online together with this communication. \\IBVSfig{21cm}{5813_f1.ps}{All H$\\alpha$ profiles measure from late 2000 to early 2006. The vertical offset of the profiles is proportional to time and corresponds to 25 days per continuum unit. The lowermost spectra date from Nov.\\ 1, 2000 (left), Sep.\\ 9, 2002 (middle) and Aug.\\ 23, 2004 (right), respectively.} ", "conclusions": "" }, "0801/0801.0038_arXiv.txt": { "abstract": "$JHK_s$ near-infrared photometry of stars in the Phoenix dwarf galaxy is presented and discussed. Combining these data with the optical photometry of Massey et al. allows a rather clean separation of field stars from Phoenix members. The discovery of a Mira variable (P = 425 days), which is almost certainly a carbon star, leads to an estimate of the distance modulus of $23.10\\pm 0.18$ that is consistent with other estimates and indicates the existence of a significant population of age $\\sim 2$ Gyr. The two carbon stars of Da Costa have $M_{bol} = -3.8$ and are consistent with belonging to a population of similar age; some other possible members of such a population are identified. A Da Costa non-carbon star is $\\Delta K_s \\sim 0.3$ mag brighter than these two carbon stars. It may be an AGB star of the dominant old population. The nature of other stars lying close to it in the $K_s,(J-K_s)$ diagram needs studying. ", "introduction": "The present investigation of the Phoenix dwarf galaxy is part of a programme to study local group galaxies using the Japanese - South African 1.4m Infrared Survey Facility (IRSF) and {\\sc sirius} three-channel camera (Nagashima et al. 1999, Nagayama et al. 2003) at SAAO Sutherland. Phoenix is a member of the Local Group and the most distant of the Milky Way's satellite galaxies (e.g. Grebel (1999) fig. 3). It was discovered by Schuster \\& West (1976) who originally suggested it might be a globular cluster; Canterna \\& Flower (1977) established that it was a galaxy. Though its overall properties are consistent with a classification as a dwarf spheroidal, it also contains a relatively small young component and is thus often referred to as a dIrr/dSph (e.g. Mateo 1998). It is associated with an off-centre H{\\sc i} cloud (Oosterloo, Da Costa \\& Staveley-Smith 1996; Young \\& Lo 1997; St-Germain et al. 1999). The origin of this cloud is not clear, although it may be formed from supernovae winds associated with the most recent epoch of star formation in the galaxy (Young et al. 2007). Though there have been a number of optical studies of Phoenix, this seems to be the first to describe $JHK_s$ observations. ", "conclusions": "By combining our own $JHK_s$ observations with the optical photometry of Massey et al. (2007) it has been possible to make a rather clean separation of Phoenix members from field stars. A clear RGB of an old population is found together with a few highly evolved stars. A Mira variable, almost certainly a carbon star, with a period of 425 days is present in the galaxy and leads to an estimate of $23.10 \\pm 0.18$ for the distance modulus in agreement with other estimates. The kinematics of carbon Miras in our Galaxy (Feast et al. 2006) suggest an age of $\\sim 2\\;$Gyr for this star. Since Miras are relatively short-lived objects this implies a significant population of this age. The two Da Costa carbon stars have $M_{K} = -6.2$ or $M_{bol} = -3.8$ (based on an estimate of the bolometric correction from the work of Frogel et al.(1980)). These luminosities are consistent with an age $\\sim 1$ to a few Gyr (see e.g. the luminosities of carbon stars in LMC clusters (Frogel et al. 1990)). Whilst most of the stars fainter than $K_s \\sim 17.2$ are found to be members of an old RGB population, a significant number of them are identified as probably AGB stars of intermediate age. They are likely to belong to the same population as the carbon stars. In this connection, we note that a feature in the colour-magnitude diagram of Holtzman et al. (2000) (their fig. 2), starting at $V$ or $I$ of $\\sim 24.0$, $(V-I) \\sim 0$ and sloping to higher luminosities and redder colours may be a subgiant branch of intermediate age stars. It is reasonably well fitted by a 1 Gyr isochrone (z = 0.002) from Girardi et al. (2002). The status of the non-carbon star Da Costa C2 which is $\\sim 0.3$ mag brighter than the two carbon stars at $K_s$ is uncertain. It seems most likely to be an AGB star of an old population. Whether other stars of about the same luminosity and colour to C2 are also old AGB stars or carbon stars of an intermediate age population requires further spectroscopic work." }, "0801/0801.2451_arXiv.txt": { "abstract": "The rich oscillation spectra determined for the two stars, $\\nu$~Eridani and 12~Lacertae, present an interesting challenge to stellar modelling. The stars are hybrid objects showing a number of modes at frequencies typical for $\\beta$ Cep stars but also one mode at frequency typical for SPB stars. We construct seismic models of these stars considering uncertainties in opacity and element distribution. We also present estimate of the interior rotation rate and address the matter of mode excitation. We use both the OP and OPAL opacity data and find significant difference in the results. Uncertainty in these data remains a major obstacle in precise modelling of the objects and, in particular, in estimating the overshooting distance. We find evidence for significant rotation rate increase between envelope and core in the two stars. Instability of low-frequency g-modes was found in seismic models of $\\nu$~Eri built with the OP data, but at frequencies higher than those measured in the star. No such instability was found in models of 12~Lac. We do not have yet a satisfactory explanation for low frequency modes. Some enhancement of opacity in the driving zone is required but we argue that it cannot be achieved by the iron accumulation, as it has been proposed. ", "introduction": "} Transport of chemical elements is still not well understood aspect of stellar interior physics. In the case of the upper main sequence there are uncertainties regarding the extent of mixing of the nuclear reaction products beyond the convective core, known as the overshooting problem, as well as, regarding the survival of element stratification in outer layers caused by selective radiation pressure and diffusion. Closely related is another difficult problem of the angular momentum transport because there is a role of rotation in element mixing. The upper main sequence pulsators, $\\beta$ Cephei stars, are potential source of constraints on modelling the transport processes massive stars. Recently, Miglio et al. (2007b) and Montalb\\'an et al. (2007) studied sensitivity of the oscillation frequencies in B stars to the rotationally induced mixing. A comprehensive survey of properties of these variables was published by Stankov \\& Handler (2005). Pulsation encountered in $\\beta$ Cep stars frequently consists in excitation of a number of modes, which differ in their probing properties. They are found most often in evolved objects, where low order nonradial modes have mixed character: acoustic-type in the envelope and gravity-type in the deep radiative interior. Frequencies of such modes are very sensitive to the extent of overshooting. Their rotational splitting is a source of information about the deep interior rotation rate. Pulsation is found both in very slow ($<10$~km~s$^{-1}$) and very rapid ($>200$~km~s$^{-1}$) rotating stars. Thus, there is a prospect for disentangling the effect of rotation in element mixing. First limits on convective overshooting and some measures of differential rotation have been already derived from data on the $\\beta$ Cep stars HD 129929 (Aerts et al. 2003), $\\nu$~Eri (Pamyatnykh et al. 2004 (PHD), and Ausselooss et al. 2004). Unfortunately, both objects are very slow rotators. To disentangle the role of rotation in element mixing, we need corresponding constraints for more rapidly rotating objects. It seems that we understand quite well the driving effect responsible for mode excitation in $\\beta$ Cep stars. The effect arises due to the metal (mainly iron) opacity bump at temperature near 200 000 K. Yet, the seismic models of $\\nu$~Eri, which reproduce exactly the measured frequencies of the dominant modes, predict instability in a much narrower frequency range than observed. As a possible solution of this discrepancy, PHD proposed accumulation of iron in the bump zone caused by radiation pressure, following solution of the driving problem for sdB pulsators (Charpinet at al. 1996). PHD have not conducted calculations of the abundance evolution in the outer layers. Instead, they adopted an {\\it ad hoc} factor 4 iron enhancement showing that it leads to a considerable increase of the instability range but also allows to fit the frequency of the troublesome p$_2$ dipole mode. However, recent calculations of the chemical evolution made by Bourge et al. (2007) showed that the iron accumulation in the $\\beta$ Cep proceeds in a different manner than in the sdB stars. Unlike in latter objects, in photospheres of $\\beta$ Cep stars the iron abundance is significantly enhanced for as long as it is enhanced in the bump. In fact, the photospheric enhancement is always greater (P.-O. Bourge private communication). However, neither $\\nu$~Eri nor 12~Lac shows abundance anomaly in atmosphere except perhaps for the nitrogen excess, [N/O]=$0.25\\pm0.3$, in the former object (Morel et al. 2006). So there is no clear evidence for element stratification and we are facing two problems: what is the cause of element mixing in outer layers of this star, and how to explain excitation of high frequency modes ($\\nu>7$~cd$^{-1}$) and the isolated very low frequency mode at $\\nu=0.43$~cd$^{-1}$. Unlike most of $\\beta$ Cep stars, where frequencies of detected modes are confined to narrow ranges around fundamental or first overtone of radial pulsation, modes in $\\nu$~Eri and 12~Lac are found in wide frequency ranges. Both stars are hybrid objects with simultaneously excited low-order p- and g-modes (typical for $\\beta$~Cep stars) as well high-order g-modes (typical for SPB stars). The greatest challenge is to explain how the latter modes are driven. Our work focuses on these two stars, which are best studied but not the only hybrid pulsators. Other such objects are $\\gamma$~Pegasi, $\\iota$~Her, HD~13745, HD~19374 (De Cat et al. 2007). Very recently Pigulski \\& Pojma\\'nski (2008) reported a likely discovery of five additional objects. In the next section, we compare oscillation spectra and provide other basic observational data for $\\nu$~Eri and 12~Lac. Seismic models of these two objects are presented in Section 3, after a brief description our new treatment of the overshooting. In the same section, we give for both objects the estimate of rotation rate gradient in the abundance varying zone outside core. Section 5 is devoted to the problem of mode excitation. Finally, in Section 6, after summarizing the results, we discuss the matter of element mixing in outer layers and what is needed for progress in B star seismology. \\vspace{4mm} ", "conclusions": "We believe that our seismic models of $\\nu$~Eridani yield a good approximation to its internal structure and rotation but there is room for improvement and a need for a full explanation of mode driving. We did not fully succeed in the interpretation of the oscillation spectrum of $\\nu$ Eridani with our standard evolutionary models and our linear nonadiabatic treatment of stellar oscillations. The frequency misfit between the high frequency peak to the $\\ell=1, {\\rm p}_2$ mode is much reduced in models built with the OP opacity data allowing large overshooting but such models are much cooler ($2\\sigma$) than the mean colour of the star implies. Such models also nearly avoid the driving problem. Similar conclusions regarding models allowing large overshooting distance were reached by Ausseloos et al. (2004). As for the consequences of usage of the OP instead of OPAL opacity data for mode instability, our finding agrees with more general observation of Miglio et al. (2007a) that models using the former data predict wider frequency ranges. An assessment of the inward rise of the angular rate in the $\\mu$-gradient zone was made for both, $\\nu$~Eri and 12~Lac, yielding the values around five for the ratio of the core to envelope rate. The surface equatorial velocities in the two stars are very different. For the former object our seismic estimate gives about 6 km~s$^{-1}$. For 12~Lac, our estimate of about 50 km~s$^{-1}$ is less certain because data on rotational splitting are much poorer but the value agrees with estimate of Desmet (private communication), based on his analysis of line profile changes. There is a large difference in rotation rate between the two stars but unfortunately the accuracy of our modelling is insufficient for addressing the question of the relation between the overshooting distance and rotation. The surface rotation rate in $\\nu$~Eri is indeed very low and this was the reason why PHD suggested that chemical element stratification in this star may be sustained. Hovever, a closer look at the problem reveals that the effect of rotation on element distribution should not be ignored, even at the equatorial velocity of few km~s$^{-1}$. As Bourge et al. (2007) showed for their $10 M_\\odot$ model, the two-fold excess of the iron abundance in the driving zone is produced in the time scale comparable with main sequence life time. The process is slower than in less massive objects (Seaton 1999) but fast enough so that if it goes unimpeded a significant excess of iron could be created. The effect of rotation could be ignored if the time, $\\tau_{\\rm mc}$, for meridional circulation to travel from the depth where most of iron is moved up ($T\\approx4\\times10^5$K) to the bottom of the convective zone ($T\\approx2\\times10^5$K) around the iron opacity bump is longer than the star age. To estimate $\\tau_{\\rm mc}$ we use the well-known expression (see e.g. Tassoul 2000) for the speed of the meridional flow \\begin{equation} V_{\\rm mc}={\\epsilon R\\over\\tau_{\\rm KH}}f(x)P_2(\\cos\\theta), \\end{equation} where $\\epsilon={\\Omega^2R^3/GM}$, $\\tau_{\\rm KH}={GM^2\\over RL}$ is the Kelvin-Helmholtz time, and $$f(x)={4\\over3}\\left({\\partial\\ln\\rho\\over\\partial\\ln T}\\right)_p {x^5\\over1-\\nabla/\\nabla_{\\rm ad}}\\left(1-{\\Omega^2\\over2\\pi G\\rho}\\right).$$ The last term, known as the Gratton-\\\"Opik term, is kept though it is of higher order because it is important near the surface where density $\\rho$ is low, even if $\\epsilon<<1$ and otherwise linearization is valid. In our model of $\\nu$~Eri, the the Gratton-\\\"Opik term is -3.4 at the depth where most of iron is supposed to be pushed up. \\begin{equation} \\tau_{\\rm mc}={\\tau_{\\rm KH}\\over\\epsilon}{\\Delta x\\over|f|} \\end{equation} We evaluate $\\tau_{\\rm mc}$ at the layer where $T=4\\times10^5$K. and for $\\Delta x$ use the distance to the bottom of the convective zone. For the $\\nu$~Eri model, we have $\\epsilon=1.2\\times10^{-4}$, $\\tau_{\\rm KH}=5.9\\times10^4$yr, $x_b=0.907$, $\\Delta x=0.05$, $f=-15$. With this numbers, we get from Eq.\\,(6) $\\tau_{\\rm mc}=1.14\\times10^6$y, which is much less than the age of the star, $1.75\\times10^7$y according to our seismic model. Thus, even in this slowly rotating star the effect of rotation cannot be ignored. For our best model of 12~Lac, we find $\\tau_{\\rm mc}$ by about four orders less than in $\\nu$~Eri. Yet, as we may see in Fig.\\,8, the problem with driving the low frequency mode is even greater than in the case of $\\nu$~Eri. Meridional flow and/or turbulence developed through instability induced by rotation are expected to prevent accumulation of iron in the driving zone and in photosphere. This is a likely explanation of apparently normal chemical atmospheric composition of $\\beta$~Cephei stars, including such slow rotators as $\\nu$~Eri. This is also the reason to reject iron accumulation in the driving zone as the solution of the driving problem, whose solution must be searched in a different way. We attach greatest hope for solution of the driving problem with improvements in stellar opacity data. Thus, we would like to encourage further effort in this field. Reliable opacity data are essential for B star seismology. Data on mode frequencies in $\\nu$~Eri and 12~Lac are abundant and accurate enough for probing rotation and structure of the $\\mu$-gradient zone above the convective core. However, to answer open questions regarding macroscopic transport of elements and angular momentum, we need accurate microscopic input data, especially on opacity. Only if we have measurements of rotational splitting, our inference on differential rotation does not rest on very precise model of stellar structure, but such data are rare. Regarding observational work, we view as most important new spectroscopic observations of the two stars aimed at improving the value of the mean effective temperature and leading to the $\\ell$ and $m$ identifications for a greater number of modes." }, "0801/0801.0454_arXiv.txt": { "abstract": "We examine the infrared properties of 43 high redshift (0.1\\,$<$\\,z\\,$<$\\,1.2), infrared-luminous galaxies in the Extended Groth Strip (EGS), selected by a deep 70\\,$\\mu$m survey with the Multiband Imaging Photometer on \\emph{Spitzer} (MIPS). In addition and with reference to starburst-type Spectral Energy Distributions (SEDs), we derive a set of equations for estimating the total infrared luminosity ($L_{IR}$) in the range 8--1000\\,$\\mu$m using photometry from at least one MIPS band. 42 out of 43 of our sources' optical/infrared SEDs ($\\lambda_{observed}$\\,$<$\\,160 $\\mu$m) are starburst-type, with only one object displaying a prominent power-law near-infrared continuum. For a quantitative analysis, models of radiation transfer in dusty media are fit onto the infrared photometry, revealing that the majority of galaxies are represented by high extinction, A$_v$\\,$>$\\,35 and for a large fraction ($\\sim$\\,50 per cent) the SED turns over into the Rayleigh-Jeans regime at wavelengths longward of 90\\,$\\mu$m. For comparison, we also fit semi-empirical templates based on \\emph{local} galaxy data, however, these underestimate the far-infrared SED shape by a factor of at least 2 and in extreme cases up to 10 for the majority ($\\sim$\\,70 per cent) of the sources. Further investigation of SED characteristics reveals that the mid-infrared (70/24\\,$\\mu$m) continuum slope is decoupled from various galaxy properties such as the total infrared luminosity and far-infrared peak, quantified by the L$_{160}$/L$_{70}$ ratio. In view of these results, we propose that these high-redshift galaxies have different properties to their local counterparts, in the sense that large amounts of dust cause heavy obscuration and are responsible for an additional cold emissive component, appearing as a \\emph{far-infrared excess} in their SEDs. ", "introduction": "The classes of galaxies known as Luminous, Ultraluminous and Hyperluminous Infrared Galaxies (LIRGs, ULIRGs and HyLIRGs), have been a major subject of focus in observational cosmology since initial, groundbreaking studies such as those of Soifer et al. (1986) and Sanders et al. (1987). These objects, discovered in abundance by the Infrared Astronomical Satellite (\\emph{IRAS}), were found to possess an array of extreme properties, including an excessive energy output, effectively manifested as a large infrared/optical ratio (Sanders et al. 1989). It is now widely accepted that the combination of a powerful UV-photon source and the presence of substantial amounts of dust are responsible for such extreme luminosities of the order of $10^{10}$--$10^{14}$$L_{\\odot}$ (e.g. Sanders $\\&$ Mirabel 1996, Genzel et al. 1998). Determining the nature of the central energy source has been the subject of numerous subsequent studies, very often supplemented by multiwavelength data (e.g. Gregorich et al. 1995, Genzel et al. 1998, Rigopoulou et al. 1999, Klaas et al. 2001, Tacconi et al. 2002, Alonso-Herrero et al. 2005). These have revealed populations dominated by powerful starburst activity, but with a non-negligible Active Galactic Nucleus (AGN) contribution found to increase with bolometric luminosity (e.g. Veilleux et al. 1997, Tran et al. 2001, Brand et al. 2006). The processes responsible for energy production and radiation transfer in the interstellar medium of such obscured systems are primarily evaluated by studying galaxy Spectral Energy Distributions (SEDs), sometimes replacing spectroscopic diagnostics in determining the nature of the central engine. For infrared-luminous objects, these processes are greatly influenced by the composition, temperature and distribution of interstellar dust. Quantifying the infrared energy budget is, therefore, key in determining the generic properties of a galaxy and, as it is directly coupled to the central energy source, important in estimating the integrated cosmic star formation history. Cases where photometric data is scarce, benefit from various methods of data extrapolation and interpolation: construction of infrared SEDs has been a popular approach of modelling emission from dusty systems, either directly from first principles or semi-empirically, bridging together contributions from different parts of the interstellar medium (e.g. Guiderdoni et al. 1998, Calzetti et al. 2000, Rowan-Robinson 2000). Such models have enabled the examination of sources with extreme properties and pronounced infrared luminosity, either by solely considering star formation processes or including contribution from AGN. In addition, establishing correlations between monochromatic flux densities or luminosities has proven extremely advantageous and has conveniently been used as a first order approximation of the infrared energy budget, as well as a diagnostic for the physical processes in galaxies, adopted extensively in evolution studies (e.g. Franceschini et al. 2001). In this paper we investigate the properties of 43 infrared-luminous galaxies from one of the first deep \\emph{Spitzer} far-IR surveys. This sub-sample is part of a population of 178 sources detected at 70\\,$\\mu$m by the Multiband Imaging Photometer on \\emph{Spitzer} (MIPS) (hereafter the 70\\,$\\mu$m population), down to a limiting flux density of 4\\,mJy at 5$\\sigma$. In Symeonidis et al. 2007 (hereafter S07), we gave a brief overview of the sources' properties and derived Star Formation Rate (SFR) estimates. Here, we aim to gain qualitative insight into their physical properties, primarily with respect to dust, comparing them to similar objects in the local universe. For this purpose and to quantitatively characterise the sample, we fit our photometry with three starburst dust models and a set of empirical local galaxy templates. The paper is set out as follows: Section 2 introduces our selection criteria for the sub-sample, including an overview of infrared colours. Section 3 analyses and compares the model SED templates, associating the quality of fits to the properties of the sources. We investigate the infrared energy budget, with reference to SED characteristics such as the continuum slope and far-IR peak, in section 4. Finally, section 5 focuses on the derivation of a set of equations to calculate the total infrared luminosity in the range 8--1000\\,$\\mu$m, with at least one MIPS band, for starburst-type sources. Our conclusions and discussion are presented in section 6. In all subsequent calculations, we have employed the following values: $H_o=71$ kms$^{-1}$Mpc$^{-1}$, $\\Omega_M=0.3$ and $\\Omega_{\\Lambda}=0.7$. ", "conclusions": "We have examined a population of 70\\,$\\mu$m selected sources, in the region of the Extended Groth Strip, focusing this study on the infrared properties of 43 objects with available spectroscopic redshifts (0.1\\,$<$\\,z\\,$<$\\,1.2). In the last section, we derived a set of equations in order to estimate the total IR luminosity with at least one MIPS band, applicable to objects with starburst-type SEDs. We would like to emphasise that our selection criteria did not photometrically bias the redshift sample with respect to the full sample or predispose our study to a particular type of galaxy. As this is a flux limited survey (see section 2), the higher redshift objects are also more luminous in the infrared and display lower optical flux. This implies that the sources with no available spectroscopic redshift are also in the LIRG/ULIRG regime. In addition, the MIPS colours (figure \\ref{fig:mipscolours}) and SEDs (figure \\ref{fig:seds}) unanimously show that all objects are photometrically similar in the mid and far-IR and their properties diverge only in the optical and near-IR, not relevant for the work described here. By fitting 4 libraries of SED templates on our infrared photometry, we have examined the properties of the sources and estimated the infrared energy budget, revealing 12 per cent starbursts, 62 per cent LIRGs and 26 per cent ULIRGs. Evaluation of the fits showed that the SK07 models perform better at reproducing the far-IR SED; their formulation adopts a two-component configuration of dust heated in the immediate locality of massive young stars and cirrus-heated dust, essentially decoupling mid and far infrared emission. As a result, since the total infrared luminosity is not directly related to the position of the SED peak or the steepness of the mid-IR continuum slope, with the SK07 templates it is possible to represent objects with strong mid-IR emission and an additional cold component. CE01 and DH02 employ a scenario with a single radiating central source where temperature varies as a function of distance from the centre. Although this has worked well with local infrared galaxies, it completely underestimates our sources' 160\\,$\\mu$m flux, with a discrepancy of at least a factor of 2 and up to 10 for extreme cases, since it does not allow for types of systems with elevated far-IR emission. Our empirical local galaxy templates, also mostly underestimate the far-IR region, indicative of the fact that an average dust temperature of 30-50\\,K is too high. Considering all above points, we propose that this deep 70\\,$\\mu$m survey has probed high-z LIRGs and ULIRGs, which are unlike local counterparts: heavy obscuration and large amounts of cold dust potentially lower than 30\\,K, appear as a \\emph{far-infrared excess} component in the SEDs, causing a significant fraction to peak at $\\lambda>$\\,90\\,$\\mu$m and steepening the continuum slope. The existence of such objects has also been put forward by other authors; e.g. Marcillac et al. (2006) have suggested that some high-z infrared-luminous objects display far-IR properties divergent from those of local galaxies because of an additional cold dust component, even if the strength of mid-IR emission is the same. This has further implications: in such systems, the continuum slope cannot be representative of various galaxy characteristics such as dust temperature, total infrared luminosity or the nature of the central energy source. Figures \\ref{fig:lumratios} and \\ref{fig:lratiopeak} demonstrate this: the mid-IR continuum slope (quantified by the $L_{70}/L_{24}$ ratio) satisfies the IRAS warm source criterion of $f_{25}/f_{60}$ $>0.17$ for sources with reduced 60\\,$\\mu$m emission and increased $\\lambda >$\\,90\\,$\\mu$m flux, as opposed to elevated 25\\,$\\mu$m emission and it is completely dissociated from the position of the SED turnover and total infrared luminosity. As this is inconsistent with the representation of far-IR emission as a single temperature black body, it is in line with our earlier suggestion of the presence of an additional cold emissive component. It is worth noting that this 70\\,$\\mu$m sample is one of the first far-IR-selected samples reaching such low flux-density limits and additional similar surveys will be of great importance in highlighting the differences (if any) between local and high redshift infrared populations. The next infrared observatory, \\emph{Herschel}, will greatly benefit this work, extending to the sub-mm part of the spectrum and, hence, enabling a complete census of high-z infrared galaxy SEDs. Finally, we would like to mention that in order to confirm our conclusions we have planned a follow-up of the work described in this paper with submillimetre observations using the Submillimetre Common-User Bolometer Array (SCUBA 2)." }, "0801/0801.0348_arXiv.txt": { "abstract": "Observations of two H$_2$CO ($3_{03}-2_{02}$ and $3_{21}-2_{20}$) lines and continuum emission at 1.3 mm towards Sgr B2(N) and Sgr B2(M) have been carried out with the SMA. The mosaic maps of Sgr B2(N) and Sgr B2(M) in both continuum and lines show a complex distribution of dust and molecular gas in both clumps and filaments surrounding the compact star formation cores. We have observed a decelerating outflow originated from the Sgr B2(M) core, showing that both the red-shifted and blue-shifted outflow components have a common terminal velocity. This terminal velocity is 58$\\pm$2 km s$^{-1}$. It provides an excellent method in determination of the systematic velocity of the molecular cloud. The SMA observations have also shown that a large fraction of absorption against the two continuum cores is red-shifted with respect to the systematic velocities of Sgr B2(N) and Sgr B2(M), respectively, suggesting that the majority of the dense molecular gas is flowing into the two major cores where massive stars have been formed. We have solved the radiative transfer in a multi-level system with LVG approximation. The observed H$_2$CO line intensities and their ratios can be adequately fitted with this model for the most of the gas components. However, the line intensities between the higher energy level transition H$_2$CO ($3_{21}-2_{20}$) and the lower energy level transition H$_2$CO ($3_{03}-2_{02}$) is reversed in the red-shifted outflow region of Sgr B2(M), suggesting the presence of inversion in population between the ground levels in the two K ladders (K$_{-1}$= 0 and 2). The possibility of weak maser processes for the H$_2$CO emission in Sgr B2(M) is discussed. ", "introduction": "The giant molecular cloud Sgr B2, located close to the Galactic center ($\\sim$44 arcmin from Sgr A*), is a well-known massive star-forming region in our Galaxy. Sgr B2 consists of an extended envelope, a hot ring and a few compact cores (e.g. Goicoechea, Rodriguze-Fernandez \\& Cernicharo 2004). The radio continuum and recombination line observations of the compact HII regions suggest that Sgr B2(N) and Sgr~B2~(M) are the two most active star forming cores in this region (Gaume \\& Claussen 1990; Gaume et al. 1995, Mehringer et al. 1993; de Pree et al. 1995, 1996, 1998). Masers, outflows and possible rotation of the two dense cores have been revealed from observations of various molecular lines at centimeter and millimeter wavelengths (Reid et al. 1988; Gaume \\& Claussen 1990; Martin-Pintado et al. 1990; Mehringer, Goss \\& Palmer 1994; Lis et al. 1993; Kuan \\& Snyder 1996; Liu \\& Snyder 1999). In addition, previous observations have shown evidence for the two hot cores to be at different evolutionary stages and to have different molecular abundances (e.g. Vogel et al. 1987; Lis et al. 1993; Miao et al. 1995; Kuan, Mehringer \\& Snyder 1996; Liu \\& Snyder 1999). H$_2$CO pervades the interstellar medium and it has a simple chemical reaction path which has been proven to be a useful probe of physical conditions (e.g. Mangum \\& Wootten 1993). The H$_{2}$CO (1$_{10}-1_{11}$) transition at 6 cm was observed in absorption against discrete continuum sources towards Sgr B2 complex with an angular resolution of $\\sim10^{\\prime\\prime}\\times20^{\\prime\\prime}$, showing nearly the same radial velocity pattern as that of the radio recombination lines (Martin-Pintado et al. 1990; Mehringer, Palmer \\& Goss 1995). These authors suggested that the H$_{2}$CO (1$_{10}-1_{11}$) transition probably arises from the surrounding gas with a relatively low mean H$_{2}$ density of $\\sim 10^{4}$ cm$^{-3}$ (Martin-Pintado et al. 1990; Mehringer, Palmer \\& Goss 1995). The millimeter H$_{2}$CO lines are an excellent tracer of H$_{2}$ density $> 10^{5}$ cm$^{-3}$ (e.g., Mangum \\& Wootten 1993). In addition, H$_{2}$CO is a planar asymmetric top molecule with very little asymmetry. The symmetry of the spin function of the molecule leads to two transition classes: ortho-H$_{2}$CO levels if the spin wavefunction is symmetric and para-H$_{2}$CO levels if antisymmetric. Since para-H$_{2}$CO is 1-3 times less abundant than ortho-H$_{2}$CO, observations of para-H$_{2}$CO have less opacity effect (Kahane et al 1984; Mangum \\& Wootten 1993). Hence, para-H$_{2}$CO appears to be a better probe to determine the physical conditions of the massive star formation regions. The millimeter/submillimeter transitions of H$_{2}$CO gas require relatively high excitation temperature and high H$_{2}$ density compared to those in the centimeter wavebands. If the brightness temperature of the continuum emission is higher than the excitation temperature, the absorption against the continuum cores can be observed in millimeter and sub-millimeter wavebands with the high angular resolution of an interferometric array (such as the Submillimeter Array,\\footnote {The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory and the Academia Sinica Institute of Astronomy and Astrophysics and is funded by the Smithsonian Institution and the Academia Sinica.} hereafter SMA). Taking advantage of the large bandwidth coverage of the SMA, we have observed multiple H$_{2}$CO lines towards Sgr B2 at 1.3 mm within a bandwidth of 2 GHz. Thus, with the same telescope system and calibration procedure, the uncertainties due to absolute flux density calibration among the different line transitions can be mitigated by measuring the line-intensity ratios, which are needed to determine physical conditions, such as kinetic temperature and H$_{2}$ number density, of the gas. In addition, the SMA is not sensitive to extended larger scale emission ($\\sim$ 50\\arcsec). Thus, the SMA observations are sensitive to the clumps of high density gas rather than the extended diffuse components. In this paper, we present the results from the SMA observations of Sgr B2 at the H$_{2}$CO lines and continuum at 1.3 mm. The paper is organized as follows: \\S 2 discusses the observations and data reduction. In \\S 3 we present the data analysis and results. In \\S 4, we present the kinematics in Sgr B2(M) by a model incorporating a spherically symmetric inflow along with a decelerating outflow. In \\S 5 we model the physical properties of the H$_{2}$CO gas in Sgr B2 using the large velocity gradient (LVG) approach. In \\S 6, we discuss the important results derived from our observations and analysis. We summarize the results in \\S7. We adopt a distance of 8 kpc to Sgr B2. ", "conclusions": "The assessment of molecular cloud mass from molecular lines can be affected by the excitation, opacity, abundance variations and gas dynamics of the molecular lines. The optically thin submillmeter dust continuum emission has been proven to be a good tracer of molecular cloud mass (Pierce-Price et al. 2000, Gordon 1995). If we take an average grain radius of 0.1 $\\mu$m and grain density of 3 g~cm$^{-3}$ and a gas to dust ratio of 100 (Hildebrand 1983, Lis, Carlstrom \\& Keene 1991), the dusty cloud mass and column density are given by the formulae (Lis, Carlstrom \\& Keene 1991) \\begin{equation} M_{H_{2}}=1.3\\times 10^{4}\\frac{e^{h{\\nu}/{\\kappa}T}-1}{Q({\\nu})} (\\frac {S_{\\nu}}{Jy})(\\frac {D}{kpc})^{2}(\\frac {{\\nu}}{GHz})^{-3}~M_{\\odot}, \\end{equation} \\begin{equation} N_{H_{2}}=8.1\\times 10^{17}\\frac{e^{h{\\nu}/{\\kappa}T}-1}{Q({\\nu}){\\Omega}}(\\frac { S_{\\nu}}{Jy})(\\frac {{\\nu}}{GHz})^{-3}~ ({\\rm cm}^{-2}), \\end{equation} \\noindent where $T$ is the mean dust temperature (K), $Q({\\nu})$ is grain emissivity at frequency ${\\nu}$, $S_{\\nu}$ is the flux density corrected for free-free emission, $\\Omega$ is the solid angle subtended by the source. Assuming $Q({\\nu})$ at 1.3 mm is 2$\\times$10$^{-5}$ and the dust temperature is 150 K for Sgr B2 ( Carlstrom \\& Vogel 1989; Lis et al. 1993; Kuan, Mehringer \\& Snyder 1996), we derived the masses, H$_{2}$ column densities and number densities. Because the continuum at 1.3 mm contains free-free emission, we estimate the physical parameters using the flux densities of the continuum corrected for free-free emission. Assuming 3.6 cm continuum emission with FWHM beam of 3$\\rlap{.}^{\\prime\\prime}8\\times 2\\rlap{.}^{\\prime\\prime}0$ (Mehringer et al. 1993) of Sgr B2(N) (K1-3) and Sgr B2(M) (F1-4) come from optically thin free-free emission ($S_{\\nu} \\varpropto {\\nu}^{-0.1}$), we estimate the free-free contribution of 4.7 and 8.4 Jy ($\\sim$ 9\\% and 24\\% of the total flux densities at 1.3 mm) at 1.3 mm towards the continuum cores K1-3 and F1-4 in our observations. Our estimates are consistent with the determinations of Lis et al. (1993) (6\\% and 33\\% of the total flux densities at 1.3 mm) and Martin-Pintado et al. (1990) ($\\sim$ 9\\% and 28\\% of the total flux densities at 1.3 mm ) for K1-3 and F1-4. The continuum flux densities at 1.3 mm corrected for free-free emission are 47.4 and 27.2 Jy for K1-3 and F1-4, respectively. From the flux densities at 1.3 cm (Gaume et al. 1995), the estimated free-free emission contributions at 1.3 mm are 0.02 Jy, 0.04 Jy and 1.26 Jy for K4, Z10.24 and MW, respectively. The corresponding continuum flux densities corrected for free-free emission are 3.5, 9.1 and 11.0 Jy, respectively. The derived H$_{2}$ masses, column densities and number densities are summarized in Table 4. The estimated H$_{2}$ masses of the Sgr B2(N) core (K1-3) and the Sgr B2(M) core (F1-4) are larger than those given by Lis et al. (1993), while the H$_{2}$ number densities are less than their results. This result is caused by the relatively larger size and higher flux densities of the continuum in our observations. The H$_{2}$CO ($3_{03}-2_{02}$) spectra show absorption against both Sgr B2(N) and Sgr B2(M) compact cores and multiple absorbing peaks. The absorptions are dominated by red-shifted gas, suggesting that the lower transition H$_{2}$CO ($3_{03}-2_{02}$) traces the cold gas in front of the continuum cores falling into the two compact cores. Previous observations showed multiple massive young stars in Sgr B2(N) and Sgr B2(M) (e.g., Gaume et al. 1995, de Pree, Goss \\& Gaume 1998). There are multiple absorbing peaks with different optical depths from our H$_{2}$CO spectra in the Sgr B2(N) and Sgr B2(M) cores, which appears to indicate that the gas is falling into the massive stars or massive star forming cores embedded at different depths in the molecular clouds (Mehringer, Palmer \\& Goss 1995). However, the angular resolution of our observations is inadequate for us to determine whether there are multiple regions present or whether the overall gravitation potential dominates the infalling gas. If infalling gas is in simple free-fall, the infalling velocities can be estimated by \\begin{equation} V_{in}=\\sqrt {\\frac {2MG}{r_{in}}}=0.09\\sqrt {\\frac {M/M_{\\odot}}{r_{in}/pc}} ~ {\\rm km~s^{-1}}, \\end{equation} \\noindent where $r_{in}$ is the infall radius, $M$ is the sum of the gas and star masses included in the $r_{in}$ and $G$ is the gravitation constant. The H$_{2}$ masses derived from the continuum are 1.4$\\times$10$^{4}$ and 7.9$\\times$10$^{3}$ M$_{\\odot}$ for the cores of Sgr B2(N) and Sgr B2(M), respectively. The VLA observations of radio continuum at 1.3 cm (Gaume et al. 1995) showed that there are three UCHII regions (K1, K2 and K3) in the core of Sgr B2(N) and four UCHII regions (F1, F2, F3 and F4) in the core of Sgr B2(M). By use of the relationships between stellar spectral type and stellar mass (Vacca, Garmany \\& Shull 1996), a total stellar mass of 68 M$_{\\odot}$ was inferred for the massive stars in the core of Sgr B2(N). The higher resolution observations (0$\\rlap{.}^{\\prime\\prime}$05) at 7 mm (de Pree, Goss \\& Gaume 1998) resolved out F1, F2, F3 and F4 into twenty-one UCHII regions, and a stellar mass of 443 M$_{\\odot}$ was inferred corresponding to the massive stars in the core of Sgr B2(M). The mass of the massive stars in the Sgr B2(M) core is six times larger than that in the Sgr B2(N) core. Taking the major axis sizes of 0.24 and 0.29 pc as the infall radii of the two cores (K1-3 and F1-4), we inferred the infall velocities of 21 and 15 km~s$^{-1}$ for Sgr B2(N) and Sgr B2(M), respectively. Hence, based on our SMA observations, we have shown that high-density molecular gas is continuously feeding onto the active star formation cores in both Sgr B2(M) and Sgr B2(N)." }, "0801/0801.1722_arXiv.txt": { "abstract": "{Inflationary scenarios based on simple non-minimal coupling $a \\phi^2 R$ and its generalizations are studied. Generalizing the form of non-minimal coupling to $ K(\\phi) R$ with an arbitrary function $K(\\phi)$, we show that the flat potential still is obtainable when $V(\\phi)/K^2(\\phi)$ is asymptotically constant. Very interestingly, if the ratio of the dimensionless self-coupling constant($\\lambda$) of the inflaton field and the non-minimal coupling constant ($a$) is small, $\\sqrt{\\lambda /a^2} \\sim 10^{-5}$, the cosmological observables for general monomial cases ($K \\sim a \\phi^m$, $V \\sim \\lambda \\phi^{2m}$) are in good agreement with recent observational data.} ", "introduction": "An early period of accelerated expansion of the universe, or inflation~\\cite{inf}, can solve many cosmological problems such as flatness problem, homogeneity problem and isotropy problem and can provide the desired initial conditions for the subsequent hot big bang cosmology~\\cite{books}. In particle physics models, inflation occurs when one or more scalar fields, the inflaton fields, dominate the energy density of the universe with their potential being overwhelming~\\cite{Lyth:1998xn}. Under such a condition, dubbed slow-roll condition, the curvature perturbation is produced which is nearly scale invariant and is heavily constrained by the measurements of the anisotropies of the CMB and the observations of the large scale structure~\\cite{obs}. The slow-roll condition says that the inflaton potential should be very flat, i.e. the effective mass of the inflaton should be very small compared with the inflationary Hubble parameter. The biggest question is the origin of the inflaton field itself. Recently Bezrukov and Shaposhnikov (BS) reported an interesting possibility that the standard model with an additional non-minimal coupling term of the Higgs field and the Ricci scalar ($\\sim a |\\phi|^2 R$) can give rise to inflation ~\\cite{Bezrukov:2007ep} without introducing any new scalar particle in the theory \\footnote{The models of chaotic inflation with nonzero $a$ were considered in various different contexts \\cite{Spokoiny:1984bd,Salopek:1988qh,Kaiser:1994vs,Komatsu:1999mt,Futamase:1987ua,Fakir:1990eg,Libanov:1998wg}.}. The authors showed that the ``physical Higgs potential'' in Einstein frame is indeed nearly flat at the large field value limit but it is also required from the COBE data $U/\\varepsilon=(0.027 \\mpl)^4$ that the ratio between the quartic coupling of the Higgs field ($\\lambda$) and the non-minimal coupling constant ($a$) should be small $\\sqrt{\\lambda/a^2}\\sim 10^{-5}$ \\footnote{One should note that when one is trying to identify the inflaton field as the Higgs field the self-coupling $\\lambda$ is of the order of unity. In that case the largeness of the non-minimal coupling is required as well.}. Here we would generalize the case of BS by taking more generic form of the nonminimal coupling and read out the required condition for the asymptotically flat potential. It is certainly worthwhile to consider the generalization since we could understand the underlying structure of the theory more closely. After reviewing the suggestion by BS in the next section, we generalize the suggestion by considering generic form of the gravity-scalar coupling term in a non-minimal way ($K(\\phi)R$) and see the general condition for getting the flat potential or the slow-roll condition in the Sec.\\ref{generalization}. In the Sec.\\ref{monomial}, we work out the monomial case with functions ($K\\sim a \\phi^m$ and $V\\sim \\lambda \\phi^{2m}$) in detail. Interestingly for any positive integer power ($m$) the scalar spectral index and the tensor-to-scalar perturbation turned out to be in good agreement with the latest cosmological observations once a combination of dimensionless self coupling constant and the coefficient of the nonminimal coupling is fixed as $\\sqrt{\\lambda_0/a_0^2}\\sim 10^{-5}$ (Details of the parameters are given in Sec.\\ref{monomial}). Summary and discussions will be followed. ", "conclusions": "In this paper, we examined the inflationary scenarios based on non-minimal coupling of a scalar field with the Ricci scalar ($\\sim K(\\phi)R$). Taking conformal transformation, the resultant scalar potential in the Einstein frame is shown to be flat at the large field limit if the condition in eq.\\ref{condition} is satisfied. This is one of the main result of this paper. This class of models gets constraints from the recent cosmological observations of the spectral index, tensor-to-scalar perturbation ratio as well as the amplitude of the potential. We explicitly considered the monomial cases $K \\sim \\phi^m$ and found that this class of models are indeed good agreement with the recent observational data: $n_S \\simeq 0.964-0.975$ and $r \\simeq 0.0007- 0.008$ for any value of $m$. In fig.\\ref{rnplot}, the predicted values for $n_S$ and $r$ are depicted. We explicitly read out the condition for fitting the observed anisotropy of the CMBR by which essentially the amplitude of the potential is determined. The condition does not look natural ($\\sqrt{\\lambda/a^2}\\sim 10^{-5}$) at the first sight but we may understand this seemingly unnatural value once we embed the theory in higher dimensional space-time. Details of higher dimensional embedding of the theory and possible solution to the smallness of $\\sqrt{\\lambda/a^2}$ will be given in separate publication \\cite{large volume}." }, "0801/0801.2930_arXiv.txt": { "abstract": "Stars at the top of the asymptotic giant branch (AGB) can exhibit maser emission from molecules like SiO, H$_2$O and OH. These masers appear in general stratified in the envelope, with the SiO masers close to the central star and the OH masers farther out in the envelope. As the star evolves to the planetary nebula (PN) phase, mass-loss stops and ionization of the envelope begins, making the masers disappear progressively. The OH masers in PNe can be present in the envelope for periods of $\\sim$1000 years but the H$_2$O masers can survive only hundreds of years. Then, H$_2$O maser emission is not expected in PNe and its detection suggests that these objects are in a very particular moment of its evolution in the transition from AGB to PNe. We discuss the unambiguous detection of H$_2$O maser emission in two planetary nebulae: K~3-35 and IRAS~17347-3139. The water-vapor masers in these PNe are tracing disk-like structures around the core and in the case of K3-35 the masers were also found at the tip of its bipolar lobes. Kinematic modeling of the H$_2$O masers in both PNe suggest the existence of a rotating and expanding disk. Both PNe exhibit a bipolar morphology and in the particular case of K~3-35 the OH masers are highly polarized close to the core in a disk-like structure. All these observational results are consistent with the models where rotation and magnetic fields have been proposed to explain the asymmetries observed in planetary nebulae. ", "introduction": "Planetary nebulae (PNe) represent the final stage of evolution for intermediate mass stars (1-8 M$_\\odot$). A PN consists of an expanding gaseous shell of highly ionized gas that has been ejected by the star during the end of the AGB evolution, and a central star that will end as a white dwarf. The study of PN is very important not only because our Sun will end as a PN but also because these objects expel heavy elements like carbon and oxygen producing an enrichment of the interstellar medium. The lifetime of a PNe is about 10$^4$ years and it is expected that there are 10$^4$-10$^5$ PNe in our galaxy, from which only $\\sim$1500 have been reported (\\cite{Kohoutek01}). PNe are characterized by exhibiting a variety of shapes (e.g. \\cite{Balick02} ). Several observational studies have revealed that most of the PNe ($\\sim$75$\\%$; \\cite{Manch04}) exhibit asymmetric morphologies that go from elliptical to very collimated outflows. One of the open questions in this field is to understand the origin of these asymmetries. \\begin{table}\\def~{\\hphantom{0}} \\begin{center} \\caption{OH and H$_2$O maser emission in young PNe} \\label{tab:young} \\begin{tabular}{lcccc}\\hline PNe & OH & H$_2$O & Morphology & References \\\\\\hline% NGC~6302 & YES & NO & bipolar & Payne et al. (1988)\\\\ OH~349.36-0.20 &YES&NO&?& Zijlstra et al. (1989)\\\\ IRAS 17207-2855& YES & NO & ? & Zijlstra et al. (1989) \\\\ PK 356+2$^\\circ$ 1 & YES & NO & ? & Zijlstra et al. (1989) \\\\ IRAS 17347-3139 & YES & YES & bipolar & de Gregorio-Monsalvo et al. (2004) \\\\ IRAS 17371-2747 & YES & NO & ? & Zijlstra et al. (1989) \\\\ IRAS 17375-2759 & YES & NO & ? & Zijlstra et al. (1989) \\\\ IRAS 17375-3000 & YES & NO & ? & Zijlstra et al. (1989) \\\\ OH0.9+1.3 & YES & NO & ? & Zijlstra et al. (1989) \\\\ IRAS 17443-2949 & YES & YES & ? & Zijlstra et al. (1989); Su\\'arez et al. (2007)\\\\ IRAS 17580-3111 & YES & YES$^a$ & ? & Zijlstra et al. (1989); Su\\'arez et al. (2007)\\\\ IRAS 18061-2505 & NO & YES$^b$ & ? & Su\\'arez et al. (2007)\\\\ IC~4997 & YES & NO & bipolar & \\cite{Tamura89} \\\\ K~3-35 & YES & YES & bipolar & Engels et al. (1989); Miranda et al. (2001)\\\\ Vy2-2 &YES&NO&shell& \\cite{Davis79} \\\\ M1-92 & YES & NO & bipolar & \\cite{lepine74} \\\\ \\hline \\end{tabular} \\begin{tabular}{l} \\noindent \\scriptsize{$^a$ Probably not a PN; $^b$ No OH maser emission (\\cite{Suarez07}).}\\\\ \\end{tabular} \\end{center} \\end{table} To answer this question we should review the formation process itself in the transition between the AGB and PNe stages. After completion of the H and He core burning, the star evolves to the AGB. At the top of the AGB phase, the star is surrounded by an expanding envelope of gas and dust (\\cite{Casw74} ). The study of the star in this evolutionary phase is better pursued with radio and infrared telescopes and the presence of neutral atomic and molecular gas in the envelope of the star is common (e.g \\cite{Mufson75}; \\cite{Huggins96}; \\cite{Josselin03}; \\cite{Bujarrabal06} ). For AGB stars with oxigen-rich envelopes it is possible to detect maser emission of one or more molecules, such as SiO, H$_2$O and OH that appear stratified in the envelope (e.g. \\cite{reidmoran81}; \\cite{Elitzur92} ). As the slow and massive mass loss of the late AGB phase stops, the gravity contracts the core heating it up, and the star enters its PN phase, from this moment the maser conditions will disappear in short timescales (\\cite{Lewis89} , \\cite{Gomez90}). In particular, the water molecules are expected to disappear in a timescale of decades, and only OH masers seem to persist for a considerable time ($\\sim$1000 yr; \\cite{Kwok93}). In this way H$_2$O masers are not expected to be present in a regular PN. However, two PNe (K3-35 and IRAS 17347-3139; \\cite{Miranda01}; \\cite{deGre04} ) have been found to harbor H$_2$O and OH maser emission, suggesting that these objects can be in an early stage of their evolution as PNe. Recently, two more PNe have been reported to exhibit H$_2$O maser emission and interferometric observations are required to confirm the association (\\cite{Suarez07}). In this talk we will present details about the kinematics and polarization of the maser emission in the two confirmed PNe: K3-35 and IRAS~17347-3139. \\begin{figure} \\includegraphics[height=5.3in,width=5.3in,angle=0]{ygomez_fig1.eps} \\caption{Left: Contour images of the radio continuum observed with the VLA in K~3-35 (\\cite{Miranda01}) and IRAS~1737-3139 (\\cite{Tafoya07b}). Right: The respective water maser spots with a disk fit (\\cite{Uscanga07}). }\\label{fig:torus} \\end{figure} ", "conclusions": "\\label{sec:concl} The presence of H$_2$O and OH maser emission in PNe is not common, and their detection suggest that these objects are in a very early phase of their evolution. In particular the presence of H$_2$O maser emission has been confirmed only in two PNe (K~3-35 and IRAS~17347-3139), and from a recent H$_2$O survey, made with the Robledo de Chavela Antenna, two more PNe seem to be detected. The H$_2$O maser positions in K~3-35 and IRAS~17347-3139 show kinematics consistent with masers located in rings. From modeling, it is derived that in K~3-35 there is a rotating and expanding disk perpendicular to the bipolar outflow. There is evidence of magnetic fields in K~3-35 that could help to constrain the models for jets and bipolar morphologies in PNe." }, "0801/0801.2876_arXiv.txt": { "abstract": "{Determining the structure of and the velocity field in prestellar cores is essential to understanding protostellar evolution.} {We have observed the dense prestellar cores L~1544 and L~183 in the $N = 1 \\rightarrow 0$ rotational transition of CN and \\thcn \\ in order to test whether CN is depleted in the high--density nuclei of these cores.} {We have used the IRAM~30~m telescope to observe along the major and minor axes of these cores. We compare these observations with the 1~mm dust emission, which serves as a proxy for the hydrogen column density.}{ We find that while CN\\jone\\ is optically thick, the distribution of \\thcn\\jone\\ intensity follows the dust emission well, implying that the CN abundance does not vary greatly with density. We derive an abundance ratio of $\\rm [CN]/[\\hh]=\\dix{-9}$ in L~183 and $1-3$\\tdix{-9} in L~1544, which, in the case of L~183, is similar to previous estimates obtained by sampling lower--density regions of the core.}{We conclude that CN is not depleted towards the high--density peaks of these cores and thus behaves like the N-containing molecules \\nnhp\\ and \\nhhh. CN is, to our knowledge, the first C--containing molecule to exhibit this characteristic.} ", "introduction": "Knowledge of the structure and kinematics of prestellar cores is important to our understanding of protostellar evolution. Cores which are close to the critical point at which collapse sets in are representative of the preliminary phase of evolution which subsequently leads to the formation of a protostar. In this context, the cores with the highest central density and/or column density are the most interesting; but the fact that CO and many other tracers of the kinematics deplete on to dust grain surfaces at densities above \\dix{5}~\\ccc\\ \\citep{tafalla2002} has been an obstacle to progress in this area. Fortunately, observations have suggested that some N--containing species, such as \\nhhh\\ and \\nnhp, remain in the gas phase at densities above \\dix{5}~\\ccc. Given that both \\nhhh\\ and \\nnhp\\ form from \\nn\\, and that the volatility of \\nn\\ is similar to that of CO \\citep{oberg2005}, it is puzzling that even these species survive. There is some evidence that they finally freeze out at densities $\\gtrsim\\dix{6}$~\\ccc\\ \\citep{bergin2002,belloche2002,pagani2007}, and, irrespective of the reason for this behaviour, it would be useful to find other tracers of the densest gas in the cores. Radicals represent a type of ``molecule'' which tends to resist depletion; this comes about because, typically, they are both formed and destroyed by neutral--neutral substitution reactions in which both formation and destruction depend in the same manner on the abundance of an atom, such as C, N, or O. An example is NO, which is believed to be formed and destroyed by reactions with atomic nitrogen. Under these conditions, the fractional abundance of NO is independent of the degree of depletion of atomic N. In a previous article \\citep[][\\, hereafter A07]{akyilmaz2007}, we attempted to test this hypothesis by observing NO in two high density cores. We found, to our discomfiture, that the hypothesis was incorrect: NO becomes depleted in the high density cores of L~1544 and L~183 similarly to many other species. The explanation of this observational result is not entirely clear, but it probably has to do with the fraction of oxygen which remains in the gas phase at high densities. The motivation for our NO study arose partly from the issues of the nitrogen chemistry, referred to earlier. NO has been considered to be the main intermediary between atomic and molecular nitrogen, which are probably the main forms of elemental nitrogen in the gas phase in molecular clouds. However, the fact that NO appears to be not present in the densest gas suggested that there may be other intermediaries between N and \\nn\\ and that a possible candidate might be CN. CN can form from hydrocarbons, such as CH, in reactions with atomic N \\citep{pineau1990} and is destroyed, also by N, forming \\nn. Thus, if elemental carbon is present in the dense gas, CN might mediate in the formation of \\nn\\ and hence in the formation of \\nhhh\\ and \\nnhp. However, the fraction of elemental carbon which is available to form hydrocarbons in the dense regions of prestellar cores is uncertain. The reason for this uncertainty is that the main reservoir of carbon in the gas phase of molecular clouds is CO, and if CO freezes out, it might be expected that there remains little carbon to form hydrocarbons and, by extension, CN. However, from an observational point of view, one knows only that CO is at least an order of magnitude underabundant in prestellar core nuclei. There are non-thermal processes \\citep[\\eg][]{leger1985} which can maintain gas phase CO at abundances of order $\\rm [CO]/[\\hh] = 10^{-6}$ at densities of order \\dix{6}~\\ccc\\ and which might suffice to produce appreciable amounts of \\cp, a precursor of CH, following the reaction of CO with cosmic--ray produced \\hep. The present Letter presents observations of CN and \\thcn\\ which demonstrate that CN is relatively abundant in the high density cores of L~183 and L~1544. In section~2, we summarize the observations and, in section~3, we present the basic observational results. Section~4 contains a brief discussion of the implications of our observational results for chemical models. ", "conclusions": "The main conclusion of this work is that CN remains in the gas phase at densities close to $106$~\\ccc\\ in prestellar cores. This result is surprising, given that CO, which is the main repository of gas--phase carbon, is clearly depleted at such densities. It follows that CN can serve as a kinematic tracer of the high--density material. The profiles of \\nnhp\\jone\\ from A07 and of \\thcn\\ from the present observations are in generally good agreement, suggesting that both molecules trace the same density regime. It is worth noting also that a double--Gaussian fit to the \\thcn\\ spectrum in L~1544 at offset $(-20,20)$ yields one component with a velocity of $7.300\\pm0.011$~\\kms\\ and a FWHM of $0.130\\pm0.025$~\\kms. The associated upper limit on the kinetic temperature is $10.0\\pm3.8$~K, which is in good agreement with the value from the best fit temperature profile of \\cite{crapsi2007} at a distance of 30\\arcsec, where $\\tkin=10$~K. In turn, this implies that the non-thermal linewidth is at most 0.65 times the thermal linewidth, and hence that turbulence has almost completely dissipated. CN is important also as a possible tracer of the magnetic field in dense cores, by means of the Zeeman effect \\citep{crutcher1999}. There exist already indirect estimates of the magnetic field in L~1544 and L~183, based upon the \\cite{chandrasekhar1953bmag} method, which are of the order of 100~$\\mu$G \\citep{crutcher2004a}. Independent and direct measurements, based on the Zeeman effect, are highly desirable. In this context, we note that OH Zeeman measurements are sensitive only to the outer core, owing to the large beam size and to the depletion of oxygen towards the centre of the core. The fact that CN is not depleted in the nuclei of L~1544 and L~183 has interesting consequences for the chemistry in these cores. One likely conclusion is that, although CO is depleted by more than an order of magnitude at densities of a few times $105$ to $106$~\\ccc, there remains sufficient CO in the gas phase to supply carbon to other, less abundant species. Quantifying this statement will require model calculations, which are under way. An expectation which is based on the model calculations is that the effective $\\rm [C]/[O]$ gas--phase abundance ratio in dense depleted regions, such as those in L~1544 and L~183, is larger than the value of $\\rm [C]/[O]=0.67$ adopted by \\cite{flower2005}. The $\\rm [CN]/[NO]$ abundance ratio is sensitive to the amount of free oxygen (i.e. not tied up in CO) in the gas phase. Our observations of L~1544 establish a lower limit to $\\rm [CN]/[NO]$ of order 0.1 in the high density region; this suggests a $\\rm [C]/[O]$ of at least 0.8. There will be consequences for HCN and HNC, and it would be useful to determine limits on the abundances of these species in the dense cores of L~1544 and L~183. We note that our suggestion (A07) that CN acts as an intermediary in the nitrogen chemistry is supported by the present results. \\begin{figure} \\def\\wa{0.55\\hsize} \\begin{center} \\includegraphics[width=0.8\\hsize,angle=-90]{pl-zero.eps} \\caption{\\thcn$(N = 1\\ra 0, F_2 = 2\\ra 1)$ spectrum at positions $(0,0)$ in L~1544 (top) and L~183 (bottom). The five observed HFS components are indicated.} \\label{fig:zero} \\end{center} \\end{figure} \\begin{figure} \\def\\wa{0.43\\hsize} \\begin{center} \\includegraphics[width=\\wa,angle=-90]{cut-l1544.eps}\\\\ \\includegraphics[width=\\wa,angle=-90]{cut-l183.eps} \\caption{Cuts through L~1544 (top) and L~183 (bottom). Filled squares: \\thcn total column density (Tables~\\ref{tab:l1544} and ~\\ref{tab:l183}) in \\cc\\ (left hand scale). Open red circles: integrated intensity (in m\\kkms: scale is indicated in top left corner) of the $(J,F_1,F)=1,0,1 \\ra 0,1,2$ transition of \\nnhp\\ (see A07). Light histogram: dust emission in MJy\\,\\unit{sr}{-1} (right hand scale).} \\label{fig:cut} \\end{center} \\end{figure}" }, "0801/0801.0274_arXiv.txt": { "abstract": "The steep source counts and negative $K$-corrections of bright submillimetre galaxies (SMGs) suggest that a significant fraction of those observed at high flux densities may be gravitationally lensed, and that the lensing objects may often lie at redshifts above 1, where clusters of galaxies are difficult to detect through other means. In this case follow-up of bright SMGs may be used to identify dense structures along the line of sight. Here we investigate the probability for SMGs to experience strong lensing, using the latest $N$-body simulations and observed source flux and redshift distributions. We find that almost all high redshift sources with a flux density above 100\\,mJy will be lensed, if they are not relatively local galaxies. We also give estimates of the fraction of sources experiencing strong lensing as a function of observed flux density. This has implications for planning follow-up observations for bright SMGs discovered in future surveys with SCUBA-2 and other instruments. The largest uncertainty in these calculations is the maximum allowed lensing amplification, which is dominated by the presently unknown spatial extent of SMGs. ", "introduction": "\\label{sec:Introduction} Sources that are picked up in deep extragalactic millimetre (mm) and submillimetre (submm) wavelength surveys are far from being ordinary galaxies. Due to strong evolution, and also driven by the typically coarse angular resolution of current surveys at these wavelengths, mm/submm sources are rare and luminous compared with well-studied populations of optical galaxies. Submm galaxies (SMGs) are usually interpreted as an early rapidly star-forming phase, perhaps driven by a major merger, in the sequence that will ultimately become a massive elliptical galaxy \\citep[e.g.,][and references therein]{Blain02}. However, gravitational lensing can also play a role, such that some fraction of SMGs may be more typical, lower-luminosity galaxies at high redshift ($z>1$) that are strongly amplified by lensing. This is already a well known effect for quasars, where the steepening source counts make lensing of increasing importance for the brightest objects \\citep[e.g.,][]{Kochanek04}. In addition, the prevalence of giant optical arcs is partly explained by the high redshifts of the sources \\citep{Wu06}. There are many anecdotal examples of SMGs that are either known or strongly suspected to be lensed. Examples include sources seen through targeted rich galaxy clusters, most spectacularly in Abell 2218 \\citep{Kneib96, Swinbank03}, and MS0451-03 \\citep{Borys04}, as well as the brightest source in the central Hubble Deep Field North, usually called HDF850.1 \\citep{Dunlop04}, the brightest source in the entire GOODS-North field, referred to as GN20 (Pope et al., submitted), and other bright well-studied sources, such as SMM\\,J14011+0252 \\citep{Ivison01, Smail05}, MIPS\\,J142824.0+352619 \\citep{Borys06}, and SMM\\,J02399-0136 \\citep{Ivison98}. It has long been recognized \\citep[e.g.,][]{Blain93} that the steep source counts and high redshifts of SMGs make the brightest ones prime lensed candidates. Despite some efforts to model the lensing contribution to the source counts \\citep[e.g.,][]{Perrotta02, Perrotta03, Negrello07} the predictions for SMG lensing are still very uncertain. The purpose of the present paper is to improve this situation, using the latest numerical models, and to try to quantify the uncertainties. In the process of this investigation we will also be able to address a related question -- namely whether bright SMGs could be used as tracers of galaxy clusters at redshifts that are high enough to be challenging for other techniques. In order to carry out this study we will need several ingredients, including the unlensed submm source counts and the redshift distribution. However, we start in section~\\ref{sec:ProbDensity} by considering the probability distribution for lensing amplification along random lines of sight. We then look at the observed and modelled source counts for SMGs in section~\\ref{sec:SourceCounts} and use these to calculate the source counts after lensing in section~\\ref{sec:Calculating}, also exploring the dependence on the redshift distribution of sources. A major simplification of our approach is to assume that the {\\it shape\\/} of the source counts is independent of redshift; in order to test this we calculate in section~\\ref{sec:evolution} lensing expectations from a specific evolutionary model that is consistent with a range of current IR and sub-mm data. Finally, we consider the uncertainties in the various components of our analysis in section~\\ref{sec:Uncertainties} and interpret our results in the context of upcoming submm surveys in section~\\ref{sec:Predictions}. ", "conclusions": "\\label{sec:Predictions} We can use our study to make approximate predictions for upcoming submm surveys. There are several relevant instruments, but we focus on the surveys which are planned with the SCUBA-2 instrument \\citep{Holland06} on the James Clerk Maxwell Telescope. There are 2 relevant $850\\,\\mu$m programmes: the SCUBA-2 Cosmology Legacy Survey, S2CLS; and the SCUBA-2 `All Sky' Survey, SASSy \\citep{Thompson07}. We have also carried out estimates for surveys at shorter wavelengths, of the sort which might be performed with the $450\\,\\mu$m array of SCUBA-2, or the SPIRE instrument on the {\\it Herschel} satellite (operating at $250$, $350$ and $500\\,\\mu$m). While it is clear that there may be many examples of strong lenses in such wide surveys, the fraction of bright sources which are lensed is significantly lower than at longer wavelengths. If one wants to find such `monsters', either as probes of line-of-sight structure or for their own intrinsic value, then one should turn to ground-based surveys in the $850\\,\\mu$m or ${\\sim}\\,1\\,$mm windows. The S2CLS plans to map approximately $20\\,{\\rm deg}^2$ to an RMS of $0.7\\,$mJy. From our number counts model we estimate that there will be about 96, 44 and 27 sources detected with $S\\,{>}\\,20$, 25 and $30\\,$mJy, respectively. The total number of sources above these flux limits which have $\\mu\\,{>}\\,2$ will be about 15 (with most of them at the high flux end) using our best estimate from the Schechter function number counts. The numbers change only slightly over this flux density range, due to the probability of lensing increasing faster than the decrease in source counts. Above $35\\,$mJy the number of expected lensed sources declines steadily. The lensed fractions are considerably smaller using our `maximal' counts model, as can be seen by referring to Fig.~\\ref{fig:mmu-combo}. There will of course always be bright sources which have negligible lensing -- hence one would like to know how bright to go before the probability of strong lensing is significant. This can be determined using Fig.~\\ref{fig:mmu-combo}. If one is prepared to accept a 1 in 3 chance of a source being strongly lensed, then one should select sources observed with $S\\,{\\ga}\\,25\\,$mJy. If one would like the chance to be 1 in 2, then that flux rises to about $30\\,$mJy. What would one do with such sources in practice? Since strong lensing is likely to come from either a galaxy cluster, or a massive galaxy, then the existence of strong lensing implies strongly clustered structure along that line of sight. Hence for each sufficiently bright source one would use follow-up observations at other wavelengths to try to establish whether strong lensing was likely, and then to see if one could find the structure which was responsible for the lensing. Multiple images or distorted morphologies of optical counterparts would be ways of determining that lensing was taking place -- these would naturally show up as part of the procedure for trying to determine counterparts in deep data at other wavelengths. For SMGs where lensing was strongly suspected, one would target the area to search for the presence of structure along the line of sight, either with X-ray, Sunyaev-Zel'dovich or `red cluster sequence' observations. The advantage of this approach is that the high redshift of the SMG sources means that it should be feasible to find cluster (or proto-cluster) lenses at higher redshifts than are easy to achieve with most other techniques. Of course, the selection effects for clusters found in this way may be complicated to quantify. Nevertheless, building up samples of $z\\,{>}\\,1$ clusters is sufficiently important for understanding structure formation (as well as constraining dark energy, etc.), that it is worth using every available method. SASSy is designed to make a shallow $850\\,\\mu$m map over approximately $4{,}000\\,{\\rm deg}^2$, or one tenth of the sky, with the possibility of extension to a larger area later. The RMS of the maps is planned to be around $30\\,$mJy, so that a robust $5\\sigma$ catalogue will have a limit around $150\\,$mJy. It may also be possible to reduce this to nearer to $100\\,$mJy using targetted repeat observations for peaks in the maps. The procedure for identifying strongly lensed sources in SASSy may be a little different than for S2CLS, and because of the brighter flux densities, the level of uncertainty in predictions of the number of lensed sources will be considerably higher. These predictions depend strongly on the counts model, the normalization of the high amplification tail of the lensing PDF, as well as the amplification cut-off imposed by the finite source size for SMGs. Since there are still huge uncertainties in all of these factors, the expectations for SASSy cover a wide range of possibilities. Using the optimistic limit of $100\\,$mJy for SASSy, the unlensed model counts give approximately 1200 sources for the SASSy catalogue. Many of these will be in the `Euclidean counts' regime, and hence should be relatively easy to eliminate as lensing candidates. These will often be already well-known galaxies, with others typically being in the {\\sl IRAS\\/}, {\\sl Akari\\/} or radio catalogues. Colours (e.g.,~$850\\,\\mu$m to radio) can be used to distinguish objects which are likely to be at higher redshift and hence have a higher likelihood of being lensed; such methods will also be necessary to eliminate Galactic clouds. Our best lensing estimate yields a total of 1000 sources in SASSy with lensing amplification $\\mu\\,{>}\\,2$, with most of them being much more strongly lensed than this limit. We expect that these extremely lensed sources will be fairly easy to distinguish from relatively nearby, intrinsically bright galaxies, and hence the chances of there being structure along the line of sight to such candidates will be very high. Follow-up of these SMGs at other wavelengths will also be easy, since they should be at least an order of magnitude brighter than the typical SCUBA sources which have been followed up in the past." }, "0801/0801.2271_arXiv.txt": { "abstract": "The post-starburst region B in M82 and its massive star cluster component have been the focus of multiple studies, with reports that there is a large population of coeval clusters of age $\\sim$1~Gyr, which were created with a Gaussian initial mass distribution. This is in disagreement with other studies of young star clusters, which invariably find a featureless power-law mass distribution. Here, we present Gemini-North optical spectra of seven star clusters in M82-B and show that their ages are all between 10 and 300~Myr (a factor of 3-100 younger than previous photometric results) and that their extinctions range between near-zero and~4~mag~($A_V$). Using new {\\it HST ACS-HRC U}-band observations we age date an additional $\\sim$30 clusters whose ages/extinctions agree well with those determined from spectroscopy. Completeness tests show that the reported `turn-over' in the luminosity/mass distributions is most likely an artefact, due to the resolved nature of the clusters. We also show that the radial velocities of the clusters are inconsistent with them belonging to a bound region. ", "introduction": "% Region B, in the local starburst galaxy M82, has been the object of debate in recent years, due to claims that it hosts a 1~Gyr-old, independent starburst region \\citep{RdG01}. Furthermore, a turn-over was reported for the cluster mass distribution. This would make it unique among young cluster populations, which are generally accepted (and theoretically expected) to have power-law distributions. As part of a larger project\\footnote{We refer the reader to \\citet[][multi-band photometry]{m82phot07} and \\citet[][spectroscopy]{isk08a} for full details of source selection, data acquisition, reduction and analysis, and full spectroscopic and photometric results.} to study the cluster population of the entire galaxy, we have utilised multi-band ({\\it UBVI}) {\\it HST-ACS} photometry of 35 Young Massive Clusters (YMCs) in the region and additionally, Gemini-North multi-object spectroscopy (GMOS-N) of seven of them, from which we constrain extinctions, ages and radial velocities. We present these results in sections \\S~2, 3 and 4 and a summary and conclusions in \\S~5. ", "conclusions": "% We have performed a detailed study of the stellar cluster population of M82 region~B, using spectroscopy and multi-band photometry, and conclude the following: \\begin{itemize} \\item The clusters are considerably younger than previously reported \\citep{RdG03a} , with ages in the range 10-300~Myr, peaking at 150~Myr. This is in agreement with the 220~Myr timescale for the last encounter with M81, the event that triggered the starburst. \\item We find significantly higher extinctions than previous studies, ranging up to $A_V\\sim2.5$~mag. However, the region has a lower overall extinction compared to the rest of the disk, hence allowing a view deep into the body of the galaxy. \\item The radial velocities of the clusters follow the galactic rotation curve, indicating that region B cannot be kinematically distinct from the rest of the disk. \\item Consideration of the detection limit for resolved clusters shows that the reported turnover in the mass/luminosity distribution of the clusters is caused by incompleteness effects (discussed in Figure.~\\ref{plot:det-limit}). \\end{itemize} Overall, our findings contradict previous claims that present region B as having formed $\\sim1$~Gyr ago, during an off-set, independent starburst episode. Our data strongly suggest that region B simply represents a line of sight into the galaxy, allowing a clear view of the cluster population of the disk, which undoubtedly formed during a galaxy-wide starburst. The results presented here form part of a large spectroscopic study of the star cluster population of M82, which will be presented in \\citet[][spectroscopy of 61 clusters]{isk08b}." }, "0801/0801.1272_arXiv.txt": { "abstract": "{Phase self-calibration (or {\\em selfcal}) is an algorithm often used in the calibration of interferometric observations in astronomy. Although a powerful tool, this algorithm presents strong limitations when applied to data with a low signal-to-noise ratio. We analyze the artifacts that the phase selfcal algorithm produces when applied to extremely noisy data. We show how the phase selfcal may generate a spurious source in the sky from a distribution of completely random visibilities. This spurious source is indistinguishable from a real one. We numerically and analytically compute the relationship between the maximum spurious flux density generated by selfcal from noise and the particulars of the interferometric observations. Finally, we present two simple tests that can be applied to interferometric data for checking whether a source detection is real or whether the source is an artifact of the phase self-calibration algorithm.} ", "introduction": "Phase self-calibration (or {\\em selfcal}) is an algorithm often used in the calibration of radio astronomical data. It was introduced by Readhead \\& Wilkinson (\\cite{Readhead1978}) and Cotton (\\cite{Cotton1979}), and it has been essential for the success of Very Long Baseline Interferometry (VLBI) imaging. Also, the antenna-based calibrations obtained from the {\\em Global Fringe Fitting} algorithm (Schwab \\& Cotton \\cite{Schwab1983}) are equivalent to a phase self-calibration. The phase selfcal will also be an algorithm widely used with future interferometric instruments, such as the Atacama Large Millimeter Array (ALMA) or the Square Kilometre Array (SKA), now under construction or planned. Optical interferometric observations (like those in the Very Large Telescope Interferometry, VLTI) will also eventually benefit from some form of selfcal, although closure phases and amplitudes are measured in optical interferometry in a very different way than in radio. Thus, the statistical analysis presented here may need some substantial changes to rigorously describe the probability of false detections by optical interferometers. Given that part of the interferometric observations obtained from all those instruments may come from very faint sources, it is important to take into account the undesired and uncontrollable effects that the instrumentation and/or the calibration and analysis algorithms applied to the data could introduce in the interferometric observations. A deep study of all our analysis tools and their effects on noisy data is essential for discerning the reliability of detections of very faint sources. Some discoveries made by pushing the interferometric instruments to their sensitivity limits could turn out to be the result of artifacts produced by the analysis tools. The main limitations of the phase self-calibration algorithm have been analyzed in many publications (e.g., Linfield \\cite{Linfield1986}, Wilkinson et al. \\cite{Wilkinson1988}). It is well known that an unwise use of selfcal can lead to imperfect images, even to the generation of spurious source components, elimination of real components, and deformation of the structure of extended sources. In this paper, we focus on the effects that phase self-calibration produces when applied to pure noise. We show that selfcal can generate a spurious source from pure noise, with a relatively high flux density compared to the rms of the visibility amplitudes. We analytically and numerically study how the recoverable flux density of such a spurious source depends on the details of the observations (the sensitivity of the interferometer, the number of antennas, and the averaging time of the selfcal solutions). Finally, we study the effects of phase self-calibration applied to the visibilities resulting from observations of real faint sources, instead of pure noise. We present two simple tests that can be applied to real data in order to check whether a given faint source is real or not, and apply these tests to real data, corresponding to VLBI observations of the radio supernova SN\\,2004et (Mart\\'i-Vidal et al. \\cite{MartiVidal2007}). ", "conclusions": "We have analyzed the consequence of the phase-self-calibration algorithm when it is applied to extremely noisy data. We have studied how this algorithm and the statistical fluctuations of the visibility phases can create a spurious source from pure noise. The flux density of the spurious source can be a considerable fraction of the rms of the visibility amplitudes. The application of other other antenna-based calibration algorithms (like the Global Fringe Fitting) to noisy data can have similar consequences to those of selfcal if the SNR cutoff of the gain solutions is set to small values. We have considered numerical and analytic studies to show how the flux density of a spurious source created by selfcal depends on the number of antennas, the sensitivity of the array, and the averaging time of the selfcal solutions. We have also presented two simple tests that can be applied to real data in order to check if the detection of a faint source could be the result of the application of an antenna-based calibration algorithm to noisy data. These tests basically relate the averaging time of the selfcal solutions and the characteristics of the closure phase distribution to the flux density of a compact source possibly present in the data. To show a practical case, we have applied these tests to a set of real VLBI observations of supernova SN\\,2004et and found good agreement between the flux density recovered by CLEAN from the (phase-referenced) visibilities of this supernova (Mart\\'i-Vidal et al. \\cite{MartiVidal2007}) and the flux density estimate provided by our reliability tests." }, "0801/0801.4661_arXiv.txt": { "abstract": "We derive fundamental parameters of the embedded cluster DBSB\\,48 in the southern nebula Hoffleit\\,18 and the very young open cluster Trumpler\\,14, by means of deep JH\\ks\\ infrared photometry. We build colour-magnitude and colour-colour diagrams to derive reddening and age, based on main sequence and pre-main sequence distributions. Radial stellar density profiles are used to study cluster structure and guide photometric diagram extractions. Field-star decontamination is applied to uncover the intrinsic cluster sequences in the diagrams. Ages are inferred from K-excess fractions. A prominent pre-main-sequence population is present in DBSB\\,48, and the K-excess fraction $f_K=55\\pm6\\%$ gives an age of $1.1\\pm0.5$\\,Myr. A mean reddening of $A_{K_s}=0.9\\pm0.03$ was found, corresponding to $A_V=8.2\\pm0.3$. The cluster CMD is consistent with the far kinematic distance of 5\\,kpc for Hoffleit\\,18. For Trumpler\\,14 we derived similar parameters as in previous studies in the optical, in particular an age of $1.7\\pm0.7$\\,Myr. The fraction of stars with infrared excess in Trumpler\\,14 is $f_K=28\\pm4\\%$. Despite the young ages, both clusters are described by a King profile with core radii $\\rc=0.46\\pm0.05$\\,pc and $\\rc=0.35\\pm0.04$\\,pc, respectively for DBSB\\,48 and Trumpler\\,14. Such cores are smaller than those of typical open clusters. Small cores are probably related to the cluster formation and/or parent molecular cloud fragmentation. In DBSB\\,48, the magnitude extent of the upper main sequence is $\\Delta\\,\\ks\\approx2$\\,mag, while in Trumpler\\,14 it is $\\Delta\\,\\ks\\approx5$\\,mag, consistent with the estimated ages. ", "introduction": "\\label{Intro} Infrared clusters represent a new class of objects, virtually undetectable before the 90's (e.g. \\citealt{Dehar97}; \\citealt{Hodapp94}). Until recently, the number of known infrared clusters and stellar groups amounted to 276, as shown in the compilation by \\citet{BiDuBa03}. A systematic survey by \\citet{DuBiSoBa03} and \\citet{BiDuSoBa03} in directions of nebulae using 2MASS\\footnote{The Two Micron All Sky Survey, All Sky data release \\citep{Skru97}, available at {\\em http://www.ipac.caltech.edu/2mass/releases/allsky/}} revealed 346 new infrared embedded clusters and candidates. Thus, the systematic study of embedded clusters is fundamental to understand their nature, to probe the physical conditions associated to the early stages of star clusters, and to derive their fundamental parameters. Just to mention a few efforts in that direction, Serpens \\citep{OT02}, NGC\\,1333 \\citep{Warin96}, NGC\\,3576 \\citep{Persi94}, AFGL\\,5142 \\citep{CNSS93}, and S\\,106 \\citep{HodRay91}. Embedded clusters may provide the clues to better understand the formation and evolution processes of star clusters and their interaction with the parent molecular cloud. Colour-colour diagrams (2-CDs) allow identification of Pre-Main-Sequence (PMS) stars and can be used to distinguish them from Main Sequence (MS) and field stars (\\citealt{CNSS93}; \\citealt{LL03}). Together with colour-magnitude diagrams (CMDs) they can be used to derive reddening values, reddening distribution, distance from the Sun and age. K-band infrared excesses originate mostly in dust envelopes and/or discs of PMS stars, and indicate their evolutionary stage up to $\\sim10$\\,Myr (\\citealt{Greaves05}; \\citealt{MB05}), or more \\citep{BoBiOrBa06}. K, and particularly L, excesses are sensitive to the presence of protoplanetary discs (\\citealt{HLL01a}; \\citealt{OJL04}). Thus, these indicators allow dating very young star clusters (e.g. \\citealt{LAL96}; \\citealt{SoBi03}). In this context, it is important to increase the number of embedded clusters studied in detail in order to establish star-formation age spreads and constrain survival time-scales of dust envelopes and circumstellar discs. Deriving locations of embedded clusters in the Galaxy by means of photometric and spectroscopic methods is useful, since these estimators might provide distances to be compared with the available kinematic ones of the nebulae (\\citealt{GG76}; \\citealt{BFS82}). Embedded clusters are typically observed up to $\\sim4$\\,kpc from the Sun \\citep{LL03}. In turn, these studies can be used to build the Galactic structure around the Sun, in particular to trace spiral arms. Improvement in distance determinations for clusters in all disc directions would contribute to the derivation of more reliable parameters of the rotation curve of the Galaxy. The rotation curve and spiral structure has been studied by \\citet{Rus03}. In this paper fundamental cluster parameters are derived using infrared photometry of the embedded cluster DBSB\\,48 and the young open cluster Trumpler\\,14. Hoffleit\\,18 is a southern nebula discovered by \\citet{Hof53}. The cluster embedded in this nebula is Dutra, Bica, Soares, Barbuy\\,48 - DBSB\\,48 - \\citep{DuBiSoBa03}, located at J2000 $\\alpha=10^h31^m29^s$, $\\delta=-58^\\circ02\\arcmin01\\arcsec$ ($\\ell=285.26^{\\circ}$, $b=-0.05^{\\circ}$). The estimated diameter of the cluster is 1.5\\arcmin. The nebula Hoffleit\\,18 was also identified as a radio H\\,II region designated G285.3+0.0 or G285.253-0.05. The derived radio velocity is $-2\\rm\\,km\\,s^{-1}$, implying in this direction a near distance of 0.3 kpc and a far distance of 5.0 kpc. Given that the object is faint, the far distance should be the correct one, placing it in the Sagittarius-Carina arm \\citep{CaHa87}. In the present paper we check the consistency of the CMD loci with the kinematic distance estimate. Trumpler\\,14 is usually classified as a young open cluster (\\citealt{VBF96}, and references therein). Besides, it is embedded in the nebula NGC\\,3372 ($\\eta\\,\\rm Car\\ Nebula$) and its optical populous nature and environment properties make it an ideal object to be compared to a bona-fide embedded cluster such as DBSB\\,48. Trumpler\\,14 contains about 13 O stars, and is a relatively massive cluster, with a mass estimated to be around $2000\\,\\ms$ \\citep{VBF96}. In Sect.~\\ref{ObsDR} the observations are described. In Sect.~\\ref{DBSB48} the structure, CMD and 2-CDs of DBSB\\,48 are discussed, field-star decontamination is applied to the CMDs, and fundamental parameters are derived. Trumpler\\,14 is dealt with similarly in Sect.~\\ref{Tr14}. Discussions and concluding remarks are in Sect.~\\ref{Conclu}. ", "conclusions": "\\label{Conclu} Before reaching the zero-age main sequence, stars are surrounded by optically thick material consisting of an infalling envelope and accretion disc that gradually disperse along the pre-main sequence phase. Because of disc-depleting processes such as irradiation by the central star, viscous accretion and mass loss due to outflow, the median lifetime of optically thick inner accretion discs may be as short as $\\rm2 - 3\\,Myr$, with the final stages of disc accretion lasting as long as $\\rm\\sim10^7\\,yr$ (\\citealt{H05}). In addition, stars in young open clusters appear to form along some period of time (e.g. \\citealt{Sagar95}, and references therein). It is in this context that the analysis of stellar density structure and stellar-mass distribution in young star clusters is important. In the present paper we studied DBSB\\,48, the cluster embedded in the H\\,II region Hoffleit\\,18, by means of JH\\ks\\ photometry, radial stellar density profiles and fraction of K-excess emission stars. Its properties were compared to those of the young open cluster Trumpler\\,14. Besides an important population of PMS stars of different ages, field-star decontamination shows that the MS has developed only the upper mass range, as expected from PMS contraction time-scales. The relatively short MS extent ($\\Delta\\,\\ks=2$) of DBSB\\,48 reflects its younger age with respect to Trumpler\\,14 ($\\Delta\\,\\ks=5$). The upper MS extent appears to be an age indicator for well populated embedded clusters. The loci of PMS stars in DBSB\\,48 are described by tracks with ages in the range $0.3-4$\\,Myr set in the CMD with reddening $A_V=8.2\\pm0.3$ and absolute distance modulus $(m-M)_{\\circ}=13.48\\pm0.3$. With a fraction of K-excess emission of $f_K=55\\pm6\\%$ the age of DBSB\\,48 results $1.1\\pm0.5$\\,Myr. Its radial density profile is well represented by a King profile with a core radius $\\rc=0.46\\pm0.05$\\,pc. For Trumpler\\,14 we derived $\\rc=0.35\\pm0.04$\\,pc. PMS stars are described by $0.1-2$\\,Myr tracks, consistent with the $1.7\\pm0.7$\\,Myr age implied by $f_K=28\\pm4\\%$. The PMS age spread suggests that star formation in both clusters did not occur as an instantaneous event, instead it lasted a time equivalent to about the cluster ages. The core radii of DBSB\\,48 and Trumpler\\,14 are similar to that of the embedded open cluster NGC\\,6611 and significantly smaller than those of classical open clusters (e.g. \\citealt{BoBi05}; \\citealt{OldOCs}). This suggests that core formation is a process partly primordial, probably associated with parent molecular cloud fragmentation, and partly related to subsequent internal dynamical evolution." }, "0801/0801.4382_arXiv.txt": { "abstract": "We are conducting a search for supermassive black holes (SMBHs) with masses below $\\sim\\! 10^7\\,\\msun$ by looking for signs of extremely low-level nuclear activity in nearby galaxies that are not known to be AGNs. Our survey has the following characteristics: (a) X-ray selection using the \\cxo\\ X-ray Observatory, since x-rays are a ubiquitous feature of AGNs; (b) Emphasis on late-type spiral and dwarf galaxies, as the galaxies most likely to have low-mass SMBHs; (c) Use of multiwavelength data to verify the source is an AGN; and (d) Use of the highest angular resolution available for observations in x-rays and other bands, to separate nuclear from off-nuclear sources and to minimize contamination by host galaxy light. Here we show the feasibility of this technique to find AGNs by applying it to six nearby, face-on spiral galaxies (NGC 3169, NGC 3184, NGC 4102, NGC 4647, NGC 4713, NGC 5457) for which data already exist in the \\cxo\\ archive. All six show nuclear x-ray sources. The data as they exist at present are ambiguous regarding the nature of the nuclear x-ray sources in NGC 4713 and NGC 4647. We conclude, in accord with previous studies, that NGC 3169 and NGC 4102 are almost certainly AGNs. Most interestingly, a strong argument can be made that NGC 3184 and NGC 5457, both of type Scd, host AGNs. ", "introduction": "The past decade has seen extraordinary improvement in our understanding of supermassive black holes (SMBHs) --- their growth and evolution, and their links with their host galaxies. We now realize that galaxies hosting SMBHs at their centers are the rule rather than the exception. The observed correlations of the masses $\\mbh$ of the SMBHs with properties of their host galaxies, for example with the bulge stellar velocity dispersion \\citep{fm00,gea00}, show that there is a close link between the formation and evolution of galaxies and of the SMBHs they host. Furthermore, a comparison of the SMBH mass function required to explain the observed luminosity function of active galactic nuclei (AGNs) with estimates of the local SMBH mass function \\citep[e.g.][]{mea04,sea04} shows that not only must SMBHs be very common in massive galaxies but that most, if not all, of these black holes are relics of AGNs active in previous epochs. Thus knowledge of the local SMBH mass function enables us to put constraints on theories of galaxy and SMBH formation and growth, and AGN lifetimes. Estimates of the local SMBH mass function are anchored by resolved stellar or gas dynamical measurements of the masses of $\\sim 30$ SMBHs \\citep[see, e.g., the review by][]{ff05}, and otherwise based on the distribution of host galaxy properties (luminosity of the bulge or bulge stellar velocity dispersion $\\sigma$) and known scaling relationships between the mass of the SMBH and these properties (most prominently $\\mbh - \\sigma$ and $\\mbh - M_\\mathrm{bulge}$ ; but see \\citealt{gh07b} for a more direct estimate ). Most of the measured SMBH masses are $\\sim 10^8 \\msun$ or greater, however, as the sphere of influence of a less massive SMBH is extremely hard to resolve even at moderate distances, even with \\hst. For example, the sphere of influence of a $10^6 \\msun$ SMBH at 15 Mpc is $\\sim 30$ milliarcseconds (mas). As a result, while different estimates of the mass function agree for $10^8 \\msun \\lesssim \\mbh \\lesssim 10^9 \\msun$ , the low-mass end ($\\mbh \\lesssim 10^6 \\msun$) often has discrepancies \\citep[see, e.g., Fig.~7 of ][ for a comparison of different authors' estimates of the mass function]{gdea07}. A second source of uncertainty at the low-mass end is the fact that it is unknown how the scaling relationships extrapolate to very late-type spiral galaxies, which have little or no bulge component, and to very low mass galaxies (dE and dSph). Yet SMBHs \\emph{do} exist in very late-type spirals, e.g.\\ NGC 4395, a spiral galaxy of type Sdm, with $\\mbh \\sim 3\\times 10^5 \\msun$ \\citep{pea05}, and in very low mass galaxies, e.g.\\ POX 52, a dwarf galaxy, with $\\mbh \\sim 3\\times 10^5 \\msun$ \\citep{bea04}. Questions that naturally arise at this point are: Do the scaling relationships break down at low masses? What determines the mass of an SMBH: the mass of the bulge or the mass of the dark matter halo? Is there a lower bound to the local SMBH mass function? A well defined sample of low-mass SMBHs is needed to answer these questions. Since we cannot detect low-mass SMBHs by their dynamical signature, looking for them by signs of their accretion activity may be the only viable way of detecting them. This of course limits detection to the subset of low-mass black holes that are active, but this fraction can be expected to be large, for the following reason. We now understand that the ``quasar era'' is a function of luminosity, with the space density of the most luminous quasars peaking at high redshift and that of lower luminosity quasars peaking at progressively lower redshifts \\citep{fea03,hms05}. This is often referred to as the ``downsizing'' of AGN activity with cosmic epoch. Moreover, at least some models of black hole growth \\citep{mea04,m04} require anti-hierarchical growth. That is, higher mass SMBHs attain most of their mass at high redshift while lower mass SMBHs grow at progressively lower redshifts. The trends of AGN downsizing and anti-hierarchical growth, extended logically to the smallest mass SMBHs, imply that these objects were active in recent times and may still be accreting at the present epoch. The Eddington luminosity of a $10^5 \\msun$ SMBH is only $\\sim \\! 10^{43}$ \\es; low-mass SMBHs, even if accreting at high rates, will be low-luminosity AGNs (LLAGNs). The converse is not true; that is, not all LLAGNs have a low-mass SMBH. In addition to a low SMBH mass, the low luminosity of an LLAGN may be caused by a low rate or radiatively inefficient mode of accretion to an SMBH of any mass \\citep[present a study of massive SMBHs accreting at very low rates]{sfea06,sgea06}. Finally, obscuration may further lower the observed luminosity. Low-ionization nuclear emission-line region (LINER) nuclei have been studied at multiple wavelengths \\citep[e.g.][]{escm02,ssd04,dsgs05,fea06} to identify LLAGNs among them, and these studies have demonstrated that these AGNs are not necessarily the same type of object with the same physical characteristics. Since we are interested specifically in the low-mass end of the SMBH mass function, it is necessary to identify among the AGNs those that can be expected to have the smallest black holes. One approach is to use mass estimators based on the luminosity of the AGN and the width of the broad component of emission lines in its spectrum. This is the technique that was used by \\citet{gh07a} in constructing their sample of low-mass SMBHs. A second approach, and the one we use, is to look for AGN activity in galaxies (late-type spirals, dwarf galaxies) where the known host galaxy-SMBH scaling relationships predict the lowest-mass SMBHs would reside. This approach has been used by \\citet{svea07,sea08} to find candidate low-mass SMBHs. However, the scaling relationships are only statistical and cannot be used to estimate the mass of a particular SMBH; thus the second approach requires an independent estimate of the SMBH mass. A search for active low-mass SMBHs requires confirmation of the presence of an AGN in each candidate nucleus, and measurement of the mass of the SMBHs in the confirmed AGNs. The method used by \\citet{gh07a} has the virtue of effectively combining all of the above into a single step. However, as the luminosity of an AGN decreases, the optical spectrum of the galaxy nucleus becomes more and more dominated by host galaxy light, and the signature of the AGN becomes difficult to detect. Even when the optical spectrum shows no clear evidence of an AGN, however, such evidence may still be present in other wavelengths, such as x-ray and radio \\citep[e.g.][]{fea04}, and infrared \\citep[e.g.][]{dea06,svea07,sea08} \\citep[See][ for a review of nuclear activity in nearby galaxies]{h08}. For the lowest-luminosity AGNs, therefore, it is possible that a system based on optical spectra would not classify the nuclei as AGNs at all. We choose to use x-ray selection to identify candidate AGNs for the following reasons: First, x-rays can penetrate obscuring material which may be hiding the line emitting regions. Second, there are fewer sources of x-rays in a galaxy than there are of optical and UV emission and so dilution of the AGN signature by host galaxy light is less of a problem. Where the luminosity of the AGN is low to begin with, even a modest amount of absorption may result in the signal being below the optical background imposed by the galaxy. Third, even if, as expected in some theories \\citep{elb95,n00,nmm03,l03} AGNs that have luminosities or accretion rates below a cut-off value do not have broad-line regions, they should still be detectable in x-rays. X-ray observations have in fact detected AGNs in what were thought to be ``normal'' galaxies \\citep[e.g.][]{mea02}. The disadvantage, as discussed in \\S\\ref{sec:disc}, is that x-ray observations by themselves cannot always distinguish between AGNs and other x-ray sources, such as x-ray binaries (XRBs) and ultraluminous x-ray sources (ULXs). Multi-wavelength data are needed to determine the type of source. As the first step towards assembling a sample of low-mass SMBHs, we are conducting an x-ray survey of nearby (within 20 Mpc), \\emph{quiescent} spiral galaxies using the \\cxo\\ X-ray Observatory to look for nuclear x-ray sources, with an emphasis on late-type spirals. The high angular resolution of \\cxo\\ is necessary to ensure that any detected source is really at the center and is not an off-nuclear source. The survey is sensitive to an (unobscured) SMBH of mass $\\mbh = 10^4 \\msun$ radiating at $\\sim\\! 2\\times 10^{-3}\\, \\LEdd$ out to the survey limit. We will present details of the survey sample selection and the \\cxo\\ observations in a future paper. Here, we present the methods used and the feasibility of detecting AGNs using these methods by applying them to six galaxies that meet selection criteria similar to those used for the survey sample, and for which x-ray data already exist in the \\cxo\\ archive. The paper is organized as follows: \\S\\ref{sec:sampsel} describes the criteria used to select the six galaxies presented here; \\S\\ref{sec:datan} describes the observations and data analysis, with individual targets discussed in \\S\\S\\ref{sec:n3169}--\\ref{sec:n5457}; the results are discussed in \\S\\ref{sec:disc}. ", "conclusions": "} The motivation for this paper was to evaluate the feasibility of detecting low-mass SMBHs in late-type spiral galaxies, which may still be accreting at the current epoch and if so should be detectable in x-rays. The six galaxies studied in this paper are not all late-type, but span the range Sa--Sd. NGC 3169 and NGC 4102 were regarded as low luminosity AGNs. None of the remaining four nuclei, however, was known to have an accreting SMBH, of any mass. NGC 3169 and NGC 4102 are of type Sa and Sb, which have massive bulges and are expected to have massive SMBHs. For galaxies of types Scd and Sd, on the other hand, the lack of a luminous AGN could mean either that there is no SMBH or that the mass of the SMBH is low. As such, the observations studied here examine both aspects of a search for accreting low-mass SMBH: First, are these objects really detectable, given that the accretion rate is expected to be low? Second, are any sources detected in the very latest type spiral galaxies that have small or no bulges? We first note that of the six galaxies presented here, all six show nuclear x-ray sources. This implies that it is a very common occurrence. In a survey of late-type spiral galaxies such as our ongoing \\cxo\\ survey, therefore, the predominant concern is not going to be detection efficiency, but rather identification of the AGNs among the detected sources. Given that the sample presented in this paper consists of only six galaxies, we do not draw statistical conclusions here of the prevalence of very low-luminosity AGNs in nearby galaxies. But we note that, as shown in \\S\\S 3.1--3.6, of the six nuclear x-ray sources, NGC 3169 is almost certainly an AGN, and NGC 4102, NGC 3184, and NGC 5457 , have very strong, though not conclusive, arguments in favor of their being AGNs. The two remaining galaxies, NGC 4713 and NGC 4647, are ambiguous but AGNs are not ruled out. We discuss below the issues such surveys will face when attempting to identify the nature of the detected sources. The diagnostic tools traditionally used to distinguish AGNs from non-AGNs \\citep[e.g.\\ optical line ratios,][]{bpt81,vo87} were developed in the course of studying luminous AGN. Dilution of the AGN emission by host galaxy light was not a serious problem and observations with low spatial resolution (several arcseconds) sufficed. In the study of AGN that are either intrinsically less luminous or are heavily obscured, however, host galaxy light becomes increasingly problematic, and surveys relying on optical spectra \\citep[e.g.][]{hfs95,gh04} require careful subtraction of the starlight using galactic spectral templates. In the weakest AGNs, however, signs of AGN emission may not be detected by the usual diagnostics. This problem will persist until optical observations with angular resolution of 1--10 mas become possible so that host galaxy light can be effectively excluded, though it can be mitigated by using regions of the spectrum where host galaxy emission is negligible, for example very high energy x-rays (tens to hundreds of keV). The detection of a compact radio source unresolved at milliarcsecond-scales, especially if the source is accompanied by jets, would also unambiguously identify the nucleus as an AGN. This has been the motivation for radio surveys like that of \\citet{nea02}. Some AGNs obscured in the optical and UV may be detectable using infrared emission line strengths and ratios \\citep[e.g.][]{dea06,svea07,sea08}. While the methods listed above allow the unambiguous identification of AGNs, observations often do not have the angular resolution or sensitivity to distinguish AGN and non-AGN flux. Other sources of radiation in the vicinity of the AGN are, for example, plasma photoionized by the AGN itself, or a nuclear star cluster. The targeted AGNs have very low luminosity and thus even moderate amounts of obscuration may cause a significant decrement in the observed flux. In most cases, therefore, the AGN contribution should not be expected to dominate the total observed flux. Consequently, identifying these AGNs requires a different approach than what can be used in the case of the more luminous AGNs (Seyferts and QSOs). AGNs can be identified using x-ray observations (e.g.\\ with \\cxo\\ and \\textit{XMM-Newton}) solely, but only if they are point sources whose inferred luminosities are greater than $\\sim\\! 10^{41}$ \\es. Below that value, AGNs can become indistinguishable from ULXs and XRBs in x-rays. ULXs can have luminosities of a few times $10^{40}$ \\es\\ \\citep[e.g.][]{sea07} and have x-ray spectra that look similar to AGN spectra. It has been suggested \\citep{srw06} that ULX spectra show a break at $\\sim\\! 5$ keV. AGNs are not known to show this break. In high quality spectra with a large number of counts it may be possible to exploit this difference to separate ULXs and AGNs. XRBs have power law spectra with $\\Gamma \\sim 2$, can show an Fe K$\\alpha$ emission line, and emit hard x-rays, all characteristics of AGNs as well. Additionally, even with \\cxo's angular resolution, the physical space probed ranges from the central $\\sim\\!$ 10 to 100 pc of the galaxy. The existence of one or more XRBs within that region would be unsurprising. However, the fact that the inferred luminosities are as high as $10^{38}$ \\es\\ severely constrains the expected number of XRBs. For example, for NGC 5457 , one of the four galaxies presented here that are not confirmed AGNs, \\citet{pea01} provide a log N-log S relation as well as the surface density of x-ray point sources as a function of radius (their Figs.~3 and 4). Approximately 12.5\\% of the sources have luminosities exceeding $10^{37}$ \\es. The surface density of sources in the innermost $0.5\\arcmin$ is $\\sim 4.75$ arcmin$^{-2}$. Therefore we may expect $\\sim\\!0.6$ sources arcmin$^{-2}$ above the luminosity cut-off in the central $0.5\\arcmin$, or $\\sim 4\\times 10^{-3}$ such sources within the \\cxo\\ source circle of radius $2.3\\arcsec$ that has been used here. NGC 5457 is of type Scd, and can be taken to be representative of the other three galaxies, NGC 4647, NGC 3184, and NGC 4713 which are types Sc, Scd, and Sd, respectively. Thus invoking XRBs and ULXs alone to explain the x-ray observations leads to physically implausible conditions, such as requiring that most, or all, quiescent, non-starburst spiral galaxies have a ULX or $\\sim \\! 100$ XRBs in the central $0.5\\arcsec-2\\arcsec$. Nevertheless, x-ray observations by themselves will usually be inadequate to distinguish AGNs from non-AGNs in any particular instance. Information from other wavebands is thus crucial for the identification process. For the six galaxies in this paper multi-wavelength photometry is summarized in Tables~\\ref{tab:lum} and \\ref{tab:n31spitz}. But here too, traditional methods of identifying AGNs using flux ratios such as $\\alpha_{OX}$, $\\alpha_{KX}$, and $f_X/f_R$, must be used with caution. The low luminosities of the AGNs mean that the emission in the two bands being compared may not be from the same object. For instance, for an obscured AGN surrounded by a nuclear star cluster, the observed x-ray flux may be from the AGN but the optical flux may be dominated by the cluster. The existence of an AGN must instead be inferred by consistency and plausibility checks considering as much of the spectral energy distribution as possible, and the goal is the rejection of the hypothesis that all of the observed properties can be explained without requiring the presence of an AGN. NGC 3184 provides a good example where different modes of observing, imaging and spectroscopy, in two wavebands, x-ray and IR, together make a compelling argument for the presence of an AGN where each observation individually is ambiguous. Nuclear star clusters are fairly common in spiral galaxies \\citep[e.g.][]{bea02,wea05,salb08}. A cluster poses two main problems. First, it makes the presence of XRBs more likely. Second, if the cluster is young and contains many O and B stars, it may overwhelm AGN emission in the UV in addition to the optical \\citep[e.g. NGC 4303,][]{cgml02}. It may be possible in some cases, as for NGC 1042 \\citep{swea08} and NGC 4102 \\citep{gvv99}, to attempt a separation of the cluster and AGN components in the optical spectrum. In addition, stellar spectra, even for late-type stars, fall faster towards the infrared than the power-laws typical of AGNs. The mid-infrared colors of AGNs, therefore, tend to be redder than those of stellar populations and this color difference can be used to infer the presence of an AGN \\citep{skea05}. In addition to the spectral energy distribution, source variability can be a discriminant, as AGNs are known to vary in all wavelengths, whereas a star cluster, say, would not. Conversely, if the variation is periodic it would rule out an AGN and argue for an XRB instead. A survey of the type discussed here finds AGNs and strong AGN candidates, but does not measure the masses of the SMBHs in those AGNs. Measurement of the mass of one of these SMBHs will be a difficult endeavor, since the sphere of influence of the black hole cannot be resolved with current technology and resources. None of the six objects studied in this paper shows broad optical emission lines whose widths could be used to estimate the BH mass, and this is likely to be typical behavior. Spectropolarimetry may uncover broad lines in polarized light in some of the AGNs. Estimates of the SMBH mass may be made by using known scaling relationships, with the caveat that the correlations are all based on SMBHs two or more orders of magnitude more massive than the ones expected to be found by the survey. The least indirect method is an application of the observed correlation between the x-ray power law slope and Eddington ratio \\citep{wmp04}. The Eddington ratio and the luminosity in turn provide an estimate of the mass of the SMBH. Other relationships that can be used are: (a) the $\\mbh$--$\\sigma$ relation \\citep{fm00,gea00}, but this method becomes more and more uncertain as the bulge itself becomes ill-defined in the latest-type spirals; (b) the $\\mbh$--$L_\\mathrm{bulge}$ relationship \\citep{kr95,md01,mh03}, which has more scatter and also faces the problem of the definition of the bulge; (c) the $\\mbh$--$v_\\mathrm{circ}$ relation \\citep{f02,bea03}, which has the advantage that it does not require the presence of a well-defined bulge; (d) the $\\mbh$--$C$ relation \\citep{gea01}, where $C$ is the concentration of light. There is also a reported relationship between black hole mass and core radio power \\citep{sea03,mwea04}, but this relationship is not as well established as the others, and is based on observations of elliptical galaxies, so its applicability to the spiral galaxies here is uncertain. Overall, there is unlikely to be one standard method of measurement that can be applied to these galaxies, but one or more of the above methods may provide useful estimates of or limits to the masses of the SMBHs in these AGNs. Mass measurements independent of the scaling relationships are possible if an object turns out to have broad emission lines, like NGC 4395, in which case line widths or reverberation mapping may be used, or if it has maser emission, like NGC 4258, in which case gas dynamics can be used. Mass measurement in other cases will have to await mas-scale angular resolution in bands other than radio to resolve the sphere of influence of these black holes. Of the six galaxies here, NGC 3169, NGC 4102, and NGC 5457 have measurements of either the stellar velocity dispersion or the luminosity of the bulge, thus allowing an estimate of their SMBH masses (assuming here that the source in M 101 is an AGN). The scatter in the $\\mbh$--$\\sigma$ and $\\mbh$--$L_\\mathrm{bulge}$ relations, together with the uncertainty in the observed flux and the bolometric correction, result in uncertainties of about an order of magnitude in the inferred Eddington ratio, but all three objects have $L/\\LEdd \\sim 10^{-4}$. This is in the range seen in low-luminosity AGNs \\citep[e.g. $L\\sim 10^{-5}\\LEdd$ for M 81;][]{yea07}, and much higher than the $L\\sim 10^{-9}\\LEdd$ seen in truly quiescent SMBHs. These observations indicate that there is indeed a population of accreting SMBHs in nearby spiral galaxies that do not show optical signs of activity but can be uncovered by looking for their x-ray emission. Such a population will answer the question of whether the bulge is the dominant component that determines the existence, and mass, of a nuclear SMBH. The discoveries of AGNs in the Sd galaxies NGC 4395 \\citep{hfsp97} and NGC 3621 \\citep{svea07}, and the strong evidence for AGNs in the Scd galaxies NGC 3184 and NGC 5457 suggest it is not, at least as far as existence is concerned. Among the SMBHs discovered in the latest-type spirals (with small or no bulges) and in the lowest-mass galaxies should be a population of SMBHs with masses less than $10^6 \\msun$, enabling a systematic study of the low-mass end of the local supermassive black hole mass function. \\begin{deluxetable}{llcccrcccr} \\tablecaption{Targets and observation parameters\\label{tab:obs}} \\tablewidth{0pt} \\tabletypesize{\\small} \\tablecolumns{10} \\tablehead{ \\colhead{Target} & \\colhead{Morph.} & \\colhead{Nucleus} & \\multicolumn{2}{c}{Coordinates (J2000)} & \\colhead{Dist.\\tablenotemark{b}} & \\colhead{Scale} & \\colhead{Obs. Date} & \\colhead{ObsID} & \\colhead{Exp.} \\\\ \\colhead{} & \\colhead {Type} & \\colhead{Type\\tablenotemark{a}} & \\colhead{RA} & \\colhead{Dec} & \\colhead{(Mpc)} & \\colhead{(pc/\\arcsec)} & \\colhead{} & \\colhead{} & \\colhead{(ks)} } \\startdata NGC 3169 & Sa & L2 & 10 14 15.0 & +03 27 58 & 19.7 & 96 & 2001 May 2 & 1614 & 2.0\\\\ NGC 3184 & Scd & \\ion{H}{2} & 10 18 17.0 & +41 25 28 & 8.7 & 42 & 2000 Jan 8 & 804 & 39.8\\\\ & & & & & & & 2000 Feb 3 & 1520 & 21.3 \\\\ NGC 4102 & Sb & \\ion{H}{2} & 12 06 23.1 & +52 42 39 & 17.0 & 82 & 2003 Apr 30 & 4014 & 4.9\\\\ NGC 4647 & Sc & \\ion{H}{2} & 12 43 32.3 & +11 34 55 & 16.8 & 81 & 2000 Apr 20 & 785 & 36.9\\\\ NGC 4713 & Sd & T2 & 12 49 57.9 & +05 18 41 & 17.9 & 87 & 2003 Jan 28 & 4019 & 4.9\\\\ NGC 5457\\tablenotemark{c} & Scd & \\ion{H}{2} & 14 03 12.6 & +54 20 57 & 7.2 & 35 & \\nodata\\tablenotemark{c} & \\nodata\\tablenotemark{c} & 750\\\\ \\enddata \\tablenotetext{a}{From \\citet{hfs97-3}. L2: Type 2 LINER, T2: Type 2 Transition object.} \\tablenotetext{b}{From \\citet{t88} except from \\citet{ssea98} for NGC 5457.} \\tablenotetext{c}{NGC 5457 (M 101) was observed multiple times. The total observation time analyzed here is given in this table. The individual observations are listed in Table~\\ref{tab:m101}} \\end{deluxetable} \\begin{deluxetable}{lrrrrrrcc} \\tablewidth{0pt} \\tablecaption{X-ray Measurements\\label{tab:det}} \\tablecolumns{9} \\tablehead{ \\colhead{Target} & \\multicolumn{6}{c}{Counts} & \\colhead{Bkg/Src} & \\colhead{HR\\tablenotemark{a}}\\\\ \\colhead{} & \\multicolumn{2}{c}{Broad} & \\multicolumn{2}{c}{Hard} & \\multicolumn{2}{c}{Soft} & \\colhead{Area Ratio} & \\colhead{}\\\\ \\colhead{} & \\colhead{Src} & \\colhead{Bkg} & \\colhead{Src} & \\colhead{Bkg} & \\colhead{Src} & \\colhead{Bkg} & \\colhead{} & \\colhead{}} \\startdata NGC 3169 & 159 & 23 & 148 & 5 & 11 & 18 & 21.4 & \\strt $+0.86^{+0.05}_{-0.03}$\\\\ NGC 3184 N & 36 & 95 & 4 & 33 & 32 & 62 & 92 & \\strt $-0.75^{+0.08}_{-0.13}$\\\\ \\phm{NGC 3184} S & 117 & 95 & 13 & 33 & 104 & 62 & 60 & \\strt $-0.77^{+0.05}_{-0.07}$\\\\ NGC 4102 all & 354 & 48 & 78 & 8 & 276 & 40 & 6.8 & \\strt $-0.55^{+0.04}_{-0.05}$\\\\ \\phm{NGC 4102} core & 171 & 48 & 68 & 8 & 103 & 40 & 52.5 & \\strt $-0.20^{+0.08}_{-0.07}$\\\\ \\phm{NGC 4102} ext & 115 & 48 & 6 & 8 & 109 & 40 & 26.1 & \\strt $-0.88^{+0.03}_{-0.06}$\\\\ NGC 4647 & 15 & 38 & 1 & 15 & 14 & 23 & 10.1 & \\strt $-0.80^{+0.04}_{-0.20}$\\\\ NGC 4713 & 10 & 4 & 1 & 0 & 9 & 4 & 21.4 & \\strt $-0.69^{+0.09}_{-0.25}$\\\\ NGC 5457 & 314 & 256 & 23 & 45 & 291 & 211 & 20.8 & $-0.86\\pm 0.03$ \\enddata \\tablenotetext{a}{Hardness ratio, HR = (H$-$S)/(H+S), where H and S are the counts in the hard and soft bands respectively, calculated using the method described in \\citet{pea06}. The tool used is available at \\url{http://hea-www.harvard.edu/AstroStat/BEHR/}.} \\end{deluxetable} \\begin{deluxetable}{rlcrcc|rcc} \\rotate \\tablewidth{0pt} \\tablecaption{NGC 5457 X-ray Measurements\\label{tab:m101}} \\tablecolumns{9} \\tablehead{ \\colhead{ObsID} & \\colhead{Obs Date} & \\colhead{Exp Time} & \\multicolumn{3}{c}{North source} & \\multicolumn{3}{c}{South source}\\\\ \\colhead{} & \\colhead{} & \\colhead{} & \\colhead{Counts} & \\colhead{Rate} & \\colhead{HR} & \\colhead{Counts} & \\colhead{Rate} & \\colhead{HR}\\\\ \\colhead{} & \\colhead{} & \\colhead{(ks)} & \\colhead{} & \\colhead{(ks$^{-1}$)} & \\colhead{} & \\colhead{} & \\colhead{(ks$^{-1}$)} & \\colhead{}} \\startdata 934 & 2000 Mar 26 & $97.8$ & 262 & $2.68$ & $-0.81^{+0.03}_{-0.04}$ & 264 & $2.70$ & \\strt $-0.70 \\pm 0.04$\\\\ 4731\\tablenotemark{a} & 2004 Jan 19 & $56.2$ & 92 & $1.64$ & $-0.64^{+0.07}_{-0.09}$ & 136 & $2.41$ & \\strt $-0.58^{+0.06}_{-0.07}$\\\\ 5300 & 2004 Mar 7 & $52.1$ & 36 & $0.69$ & $-0.75^{+0.08}_{-0.13}$ & 117 & $2.24$& \\strt $-0.60^{+0.06}_{-0.08}$\\\\ 5309 & 2004 Mar 14 & $70.8$ & 38 & $0.54$ & $-0.62^{+0.10}_{-0.14}$ & 158 & $2.24$ & \\strt $-0.64^{+0.05}_{-0.07}$\\\\ 4732 & 2004 Mar 19 & $69.8$ & 56 & $0.80$ & $-0.77^{+0.07}_{-0.09}$ & 151 & $2.17$ & \\strt $-0.58^{+0.06}_{-0.07}$\\\\ 5322 & 2004 May 3 & $64.7$ & 54 & $0.83$ & $-0.79^{+0.06}_{-0.10}$ & 144 & $2.22$ & \\strt $-0.69 \\pm 0.06$\\\\ 5323 & 2004 May 9 & $42.4$ & 42 & $0.99$ & $-0.64^{+0.09}_{-0.14}$ & 120 & $2.83$ &\\strt $-0.54^{+0.06}_{-0.08}$\\\\ 5338\\tablenotemark{a} & 2004 Jul 6 & $28.6$ & 61 & $2.14$ & $-0.56^{+0.09}_{-0.12}$ & 41 & $1.45$ & \\strt $-0.50^{+0.12}_{-0.14}$\\\\ 5339\\tablenotemark{b} & 2004 Jul 7 & $14.0$ & 18 & $1.26$ & $-0.60^{+0.13}_{-0.22}$ & 27 & $1.90$ & \\strt $-0.79^{+0.06}_{-0.15}$\\\\ 5340 & 2004 Jul 9 & $54.4$ & 81 & $1.48$ & $-0.72^{+0.06}_{-0.08}$ & 110 & $2.02$ & \\strt $-0.74^{+0.04}_{-0.07}$\\\\ 4734 & 2004 Jul 11 & $35.5$ & 31 & $0.87$ & $-0.53^{+0.13}_{-0.17}$ & 63 & $1.77$ & \\strt $-0.61^{+0.09}_{-0.11}$\\\\ 6114 & 2004 Sep 5 & $38.2$ & 91 & $2.38$ & $-0.60^{+0.08}_{-0.09}$ & 64 & $1.68$ & \\strt $-0.64^{+0.08}_{-0.11}$\\\\ 6115 & 2004 Sep 8 & $35.4$ & 96 & $2.72$ & $-0.78^{+0.05}_{-0.07}$ & 63 & $1.79$ & \\strt $-0.70^{+0.07}_{-0.10}$\\\\ 4735 & 2004 Sep 12 & $28.8$ & 70 & $2.43$ & $-0.70^{+0.07}_{-0.10}$ & 49 & $1.71$ & \\strt $-0.77^{+0.06}_{-0.11}$\\\\ 4736\\tablenotemark{a}\\tablenotemark{b} & 2004 Nov 1 & $77.4$ & 290 & $3.75$ & $-0.78^{+0.03}_{-0.04}$ & 122 & $1.58$ & \\strt $-0.67^{+0.06}_{-0.07}$\\\\ 6152\\tablenotemark{a} & 2004 Nov 7 & $26.7$ & 133 & $4.99$ & $-0.78^{+0.05}_{-0.06}$ & 51 & $1.92$ & \\strt $-0.63^{+0.09}_{-0.12}$\\\\\\enddata \\tablenotetext{a}{Source extended, or possibly image smeared.} \\tablenotetext{b}{Possible residual contamination from background flare.} \\end{deluxetable} \\begin{deluxetable}{lccccccccc} \\rotate \\tablewidth{0pt} \\tablecaption{X-Ray Spectral Fits\\label{tab:spec}} \\tablecolumns{10} \\tablehead{ \\colhead{Target} & \\colhead{Galactic} & \\colhead{Model\\tablenotemark{a}} & \\multicolumn{6}{c}{Spectral Fit Parameters} & \\colhead{$\\chi_\\nu^2$(dof)}\\\\ \\colhead{} & \\colhead{N$_{\\mathrm H}$} & \\colhead{} & \\colhead{kT} & \\colhead{N$_{\\mathrm H}$} & \\colhead{$\\Gamma$} & \\colhead{Refl.\\tablenotemark{b}} & \\colhead{Line} & \\colhead{EW} & \\colhead{}\\\\ \\colhead{} & \\colhead{($10^{20}\\, \\mathrm{cm}^{-2}$)} & \\colhead{} & \\colhead{(keV)} & \\colhead{($10^{21}\\, \\mathrm{cm}^{-2}$)} & \\colhead{} & \\colhead{} & \\colhead{(keV)} & \\colhead{(keV)} & \\colhead{} } \\startdata NGC\\,3169 & 2.86 & \\texttt{ab(pl)} & \\nodata & \\strt $99^{+46}_{-36}$ & $2.0^{+1.2}_{-1.1}$ & \\nodata & \\nodata & \\nodata & $1.1 (26)$\\tablenotemark{c}\\\\ \\phm{NGC\\,3169} & & \\texttt{ab(pl)} & \\nodata & \\strt $116^{+90}_{-52}$ & $2.6^{+2.1}_{-1.5}$ & \\nodata & \\nodata & \\nodata & $0.3 (22)$\\\\ NGC\\,4102 core & 1.79 & \\texttt{ab(pl)} & \\nodata & \\strt $0^{+1.4}$ & $2.0\\pm 0.5$ & \\nodata & \\nodata & \\nodata & $7.4(27)$\\tablenotemark{c}\\\\ & & \\texttt{ab(rn)} & \\nodata & \\strt $0^{+0.4}$ & $1.8\\pm 0.4$ & $108^{+149}_{-61}$ & \\nodata & \\nodata & $2.3 (26)$\\tablenotemark{c} \\\\ & & \\texttt{ab(rn+g)} & \\nodata & \\strt $0^{+1.9}$ & $2.3^{+0.6}_{-0.5}$ & $129$\\tablenotemark{d} & $6.4$ & $3.3$ & $0.6 (25)$\\\\ & & \\texttt{ab(rn+g)} & \\nodata & \\strt $0^{+0.6}$ & $2.2^{+0.6}_{-0.5}$ & $129^{+283}_{-85}$ & $6.4$ & $2.5$ & $1.6 (25)$\\tablenotemark{c} \\\\ \\phm{NGC\\,4102} ext & & \\texttt{ab(br)} & \\strt $1.2^{+8.6}_{-0.8}$ & $1.2^{+2.8}_{-1.2}$ & \\nodata & \\nodata & \\nodata & \\nodata & $0.5(18)$\\\\ NGC\\,5457 sep.\\ low & 1.15 & \\texttt{ab(me+pl)} & \\strt $0.3^{+0.2}_{-0.1}$ & $0^{+1.8}$ & $1.7\\pm 0.5$ & \\nodata & \\nodata & \\nodata & $0.5(28)$\\\\ \\phm{NGC\\,5457} sep.\\ high & & \\texttt{ab(pl)} & \\nodata & \\strt $1.5^{+1.1}_{-1.5}$ & $2.2^{+0.4}_{-0.3}$ & \\nodata & \\nodata & \\nodata & $0.5(40)$\\\\ \\phm{NGC\\,5457} sim.\\ low & & \\texttt{ab(me+pl)} & \\strt $0.3^{+0.3}_{-0.1}$ & $0.42^{+1.9}_{-0.42}$ & $2.0^{+0.4}_{-0.2}$ & \\nodata & \\nodata & \\nodata & $0.5 (28)$\\\\ \\phm{NGC\\,5457} sim.\\ high & & \\texttt{ab(pl)} & \\strt \\nodata & $1.1 \\pm 0.6$ & $2.0^{+0.4}_{-0.2}$ & \\nodata & \\nodata & \\nodata & $0.5 (41)$\\\\ \\phm{NGC\\,5457} merged & & \\texttt{ab(pl)} & \\strt \\nodata & $0.6 \\pm 0.4$ & $1.9 \\pm 0.2$ & \\nodata & \\nodata & \\nodata & $0.6 (70)$\\\\ \\enddata \\tablenotetext{a}{Model labels: \\texttt{ab}=\\texttt{xswabs}, photoelectric absorption; \\texttt{ga}=\\texttt{xswabs} with value frozen at Galactic column density towards this target; \\texttt{br}=\\texttt{xsbremss}, thermal bremsstrahlung; \\texttt{g}=\\texttt{gauss1d}, one-dimensional Gaussian; \\texttt{me}=\\texttt{xsmekal}, thermal plasma; \\texttt{pl}=\\texttt{powlaw1d}, one-dimensional power law; \\texttt{rn}=\\texttt{xspexrav}, power-law reflected by neutral material.} \\tablenotetext{b}{Reflection scaling factor.} \\tablenotetext{c}{Cash statistic (\\texttt{cstat} in \\textit{Sherpa}).} \\tablenotetext{d}{Unconstrained by fit.} \\end{deluxetable} \\begin{deluxetable}{cccccccc} \\tablewidth{0pt} \\tablecaption{Inferred nuclear luminosities\\label{tab:lum}} \\tablecolumns{8} \\tablehead{ \\colhead{} & \\colhead{Filter} & \\colhead{NGC 3169} & \\colhead{NGC 4102} & \\colhead{NGC 4647} & \\colhead{NGC 3184} & \\colhead{NGC 5457} & \\colhead{NGC 4713}\\\\ \\colhead{} & \\colhead{or Band} & \\colhead{(Sa)} & \\colhead{(Sb)} & \\colhead{(Sc)} & \\colhead{(Scd)} & \\colhead{(Scd)} & \\colhead{(Sd)} } \\startdata X-ray & 0.3--8 keV & 41.7 & 40.2 & 39.0\\tablenotemark{ab} & 37.3\\tablenotemark{c} & 37.5--38.5\\tablenotemark{d} & 38.6\\tablenotemark{c}\\\\ UV \\& Opt. & F300W & \\nodata & \\nodata & \\nodata & $< 39.4$ & \\nodata & \\nodata \\\\ & F336W & \\nodata & \\nodata & \\nodata & \\nodata & 39.0 & \\nodata \\\\ & F547M & \\nodata & \\nodata & \\nodata & \\nodata & 39.5 & \\nodata \\\\ & F606W & \\nodata & \\nodata & \\nodata & 39.5 & \\nodata & 40.1 \\\\ IR & $K_s$ & 42.6 & 42.8 & 38.8 & 40.9 & 41.0 & 41.0 \\\\ Radio & 15 GHz & \\nodata & \\nodata & \\nodata & \\nodata & \\nodata & $< 36.8$ \\\\ & 5 GHz & 37.2 & \\nodata & \\nodata & \\nodata & $< 35.8$ & \\nodata \\\\ & 1.4 GHz & \\nodata & 38.0 & 36.9\\tablenotemark{b} & $< 35.1$ & $< 34.9$ & $< 35.7$ \\enddata \\tablecomments{Values are $\\log(L/\\mathrm{erg\\:s}^{-1})$ in the specified bandpass in the x-ray and $\\log(\\nu L_\\nu/\\mathrm{erg\\:s}^{-1})$ where a single filter, wavelength, or frequency is given. Note that the values are derived from observations with varied instruments, PSFs, apertures, and signal-to-noise ratios, and the uncertainties can be as large as a factor of two.} \\tablenotetext{a}{0.3--12 keV.} \\tablenotetext{b}{{L}arge positional uncertainty makes it unclear whether the detected source is really the nucleus.} \\tablenotetext{c}{Based on a small number of counts. A power-law spectrum was assumed. Intrinsic absorption is unknown.} \\tablenotetext{d}{Variable source.} \\end{deluxetable} \\begin{deluxetable}{cccccccc} \\tablewidth{0pt} \\tablecaption{NGC 3184 luminosity in \\textit{Spitzer\\/} bands.\\label{tab:n31spitz}} \\tablecolumns{3} \\tablehead{ \\colhead{Band} & \\colhead{Aperture} & \\colhead{$\\log(\\nu L_\\nu)$}\\tablenotemark{a} \\\\ \\colhead{$(\\micron)$} & \\colhead{Radius} & \\colhead{(\\es)} } \\startdata 3.6 & $3\\arcsec$ & 40.6\\\\ 4.5 & $3\\arcsec$ & 40.4\\\\ 5.8 & $3\\arcsec$ & 40.7\\\\ 8.0 & $3\\arcsec$ & 40.8\\\\ 24 & $6\\arcsec$ & 41.1\\\\ 70 & $10\\arcsec$ & 41.5\\\\ 160 & $10\\arcsec$ & 41.7\\\\ \\enddata \\tablenotetext{a}{Uncertainties are 10\\% in the $3.6\\micron$, $4.5\\micron$, $5.8\\micron$, $8.0\\micron$, and $160\\micron$ bands; 4\\% at $24\\micron$; 12\\% at $70\\micron$.} \\end{deluxetable}" }, "0801/0801.1875_arXiv.txt": { "abstract": "The Einstein radius of a cluster provides a relatively model-independent measure of the mass density of a cluster within a projected radius of $\\sim 150$ kpc, large enough to be relatively unaffected by gas physics. We show that the observed Einstein radii of four well-studied massive clusters, for which reliable virial masses are measured, lie well beyond the predicted distribution of Einstein radii in the standard $\\Lambda$CDM model. Based on large samples of numerically simulated cluster-sized objects with virial masses $\\sim 10^{15}M_{\\odot}$, the predicted Einstein radii are only $15-25\\arcsec$, a factor of two below the observed Einstein radii of these four clusters. This is because the predicted mass profile is too shallow to exceed the critical surface density for lensing at a sizable projected radius. After carefully accounting for measurement errors as well as the biases inherent in the selection of clusters and the projection of mass measured by lensing, we find that the theoretical predictions are excluded at a 4-$\\sigma$ significance. Since most of the free parameters of the $\\Lambda$CDM model now rest on firm empirical ground, this discrepancy may point to an additional mechanism that promotes the collapse of clusters at an earlier time thereby enhancing their central mass density. ", "introduction": "The standard picture of the basic cosmological framework has recently come to rest firmly on detailed empirical evidence regarding the cosmological parameters, the proportions of baryonic and non-baryonic dark matter, together with the overall shape and normalization of the power spectrum \\citep[e.g.,][]{SNV06,Spergel07,BAO07}. This framework has become the standard $\\Lambda$CDM cosmological model, with the added simple assumptions that the dark matter reacts only to gravity, is initially sub-relativistic, and possesses initial density perturbations which are Gaussian distributed in amplitude. This is a very well defined and relatively simple model, with clear predictions which are amenable to examination with observations. The cooling history of baryons complicates the interpretation of dark matter on galaxy scales, especially for dwarf galaxies that traditionally have been a major focus of studies of halo structure. Clusters have the advantage that the virial temperature of the associated gas is too hot for efficient cooling, so the majority of the baryons must trace the overall gravitational potential and hence we may safely compare lensing-based cluster mass measurements to theoretical predictions that neglect gas physics and feedback. Lensing-based determinations of the mass profiles of galaxy clusters rely on detailed modeling of the strong lensing region to define the inner mass profile, and also a careful analysis of the outer weak lensing regime. The latter involves substantial corrections for instrumental and atmospheric effects \\citep{KSB}, and a clear definition of the background, free of contamination by the lensing cluster \\citep{Br05b,Medezinski}. In the center we may make use of the Einstein radius of a cluster which is often readily visible from the presence of giant arcs and provides a relatively model-independent determination of the central mass density. In the case of axial symmetry, the projected mass inside the Einstein radius $\\tE$ depends only on fundamental and cosmological constants: $M(<\\tE)=\\tE^2 (c^2/4G) D_{\\rm OL} D_{\\rm OS} / D_{\\rm LS}$, where this combination of angular diameter distances (observer-lens, observer-source, and lens-source) leads to a relatively weak dependence on the lens and source redshifts. More generally, an effective Einstein radius can be defined by axially averaging the projected surface density, which itself is well determined when there are a large number of constraints. Virtually all known massive clusters at intermediate redshifts, $0.150$.} \\label{fig:Nofz} \\end{figure} We compared the theoretical predictions with the observed $\\tE$ for four clusters, A1689, Cl0024, A1703, and RXJ1347. For the latter three we used the virial mass as given by NFW fits to the lensing observations, but for A1689 we obtained a model-independent mass directly from the lensing data, only assuming spherical symmetry (section~2.3). For each object, the predicted $\\tE$ values came up short by a factor of two compared with the observations (Figure~\\ref{fig:rEofz}). After including the measurement errors, the full probability distribution functions of the predicted Einstein radii excluded the theoretical model at 2-$\\sigma$ for each object. The total probability of the standard $\\Lambda$CDM model yielding four clusters with such large $\\tE$ is $3 \\times 10^{-5}$, a 4-$\\sigma$ discrepancy. Lensing work is now being extended to larger samples of clusters, so that in the near future we may examine more fully the relation between the Einstein radius and virial mass and its scatter, over a wider range of cluster masses. The theoretically predicted triaxiality of CDM halos implies a particular scatter in the projected concentration parameter (and thus in the Einstein radius) for a given halo mass. This scatter, which we included in our calculations, is apparently insufficient to explain the observations. If the scatter is observationally determined to be relatively small, then this would further highlight the problem we have discussed and leave the high concentrations unexplained. Determining the scatter observationally will also statistically probe the degree of triaxiality of CDM halos. To ensure the most direct, unbiased comparison, the simulated distributions should be calculated not at a fixed 3-D virial mass, but at a fixed projected, effective virial mass, defined as in section~2.3. We expect this projected virial mass to also be observationally measured in more clusters. Numerical simulations show a clear correlation between the concentration of a halo and its formation time, i.e., the time at which a significant portion of the halo mass first assembled \\citep[e.g.,][]{Neto}. This agrees with the intuitive notion that a dense halo core must have assembled at high redshift, when the cosmic density was high. Thus, the fact that observed cluster halos are apparently more centrally concentrated than is predicted in $\\Lambda$CDM suggests an additional mechanism that promotes the collapse of cluster cores at an earlier time than expected. Baryons are unlikely to help. Central cD galaxies contribute only a small fraction of the mass within the Einstein radius, which for our four clusters is $\\sim 150$ kpc enclosing a projected mass of $\\sim 2 \\times 10^{14} M_\\odot$, or a 3-D mass of $\\sim 1 \\times 10^{14} M_\\odot$. We can estimate the effect of baryons on the total mass profile using the simple model of adiabatic compression \\citep{blumenthal}. Within this model, conservation of angular momentum implies that the quantity $r M(r)$ (assuming spherical symmetry) is fixed. Assuming that we start out with a halo with the mean expected theoretical concentration (Figure~\\ref{fig:NetoMean}), the observed 3-D mass within the Einstein radius can be obtained through adiabatic compression if the enclosed baryonic mass within this radius is $\\sim 3 \\times 10^{13} M_\\odot$ in the four clusters we considered. Thus, explaining the discrepancy through adiabatic compression requires the baryonic fraction within the Einstein radius to be $\\sim 1/3$, twice the cosmic baryon fraction. This seems highly unlikely, as the observed X-ray emission yields at these radii gas fractions well below the cosmic value (e.g., see Figure~12 of \\citet{Doron} for A1689), and a cD galaxy contains only $\\sim 1 \\times 10^{12} M_\\odot$ in baryons (e.g., see \\citet{M87} for M87). Modifications in the properties of dark matter or the slope of the primordial power spectrum are generally expected to have a smaller effect on clusters than on smaller-scale objects which are predicted in $\\Lambda$CDM to have earlier formation times and higher concentrations. On the other hand, since clusters are rare objects in the standard model, primordial non-Gaussianity would significantly affect them and allow clusters to form earlier, which may also help explain other observations \\citep{silk,yoel}. Regardless of the mechanism, early collapse of cluster cores may have observable consequences if it is accompanied by star formation. Finally, we note that the fact that clusters are now being detected with masses $\\sim 10^{15} M_\\odot$ is completely consistent with the $\\Lambda$CDM model. Indeed, Figure~\\ref{fig:Nofz} shows that large numbers of clusters are expected out to significant redshifts, including $M = 2 \\times 10^{15} M_\\odot$ halos out to $z \\sim 1$, as well as more massive halos up to $\\sim 5\\times 10^{15} M_\\odot$ at lower redshift. Clearly, while large samples of halos with precise, profile-independent lensing determinations of both $\\tE$ and $\\Mv$ will make our results completely conclusive, the highly-significant discrepancy we have identified already represents a substantial challenge for $\\Lambda$CDM." }, "0801/0801.1334_arXiv.txt": { "abstract": "In the coming year, the Large Hadron Collider will begin colliding protons at energies nearly an order of magnitude beyond the current frontier. The LHC will, of course, provide unprecedented opportunities to discover new particle physics. Less well-known, however, is that the LHC may also provide insights about gravity and the early universe. I review some of these connections, focusing on the topics of dark matter and dark energy, and highlight outstanding prospects for breakthroughs at the interface of particle physics and cosmology. ", "introduction": "The Large Hadron Collider (LHC) is scheduled to begin running in the summer of 2008. Conceived around 1984 and approved in 1994, the LHC will provide the first detailed look at the weak energy scale $\\mweak \\sim 100~\\gev - 1~\\tev$ by colliding protons with protons at the center-of-mass energy $E_{\\text{CM}} = 14~\\tev$ and ultimate luminosity ${\\cal L} = 100~\\ifb/\\yr$. This is far beyond the current energy frontier, where the Tevatron collides protons and anti-protons with $E_{\\text{CM}} = 2~\\tev$ and ${\\cal L} \\sim 1~\\ifb/\\yr$. As an illustration of the power of the LHC, top quarks, discovered in 1994 with a handful of events and currently produced at the Tevatron at the rate of $\\sim 1000$ per year, will be produced at the rate of 10 Hz when the LHC reaches its design luminosity. The {\\em raison d'etre} for the LHC is the discovery of the Higgs boson and associated microphysics, including supersymmetric and other postulated particles. In recent years, however, the LHC's potential for providing insights into gravity and cosmology have taken on increasing importance. My goal here is to review some recent developments and to highlight a few scenarios in which the implications of the LHC for our understanding of gravity and the early universe may, in fact, be profound. ", "conclusions": "In the coming year, the LHC will probe the weak scale $\\mweak \\sim 100~\\gev - 1~\\tev$ in great detail. This has implications for new particle physics, but may also open up new windows on the early universe, and tests of gravity in rather unusual regimes. At present, the evidence for particle dark matter is as strong as ever. The possibility of WIMP dark matter is well-motivated, and there has been a recent proliferation of candidates. At the same time, there has also been a great deal of progress on the alternative possibility of superWIMP dark matter. In virtually all cases, the LHC will be able to produce these candidates, and in some cases, precision measurements at the LHC may be able to determine the candidate's relic density. Comparison with observations may then provide compelling evidence that particles produced at the LHC do, in fact, constitute the dark matter. Such studies will also be able to determine the microscopic properties of the WIMP particle. Colliders may also provide an interesting window on gravity in unusual environments. For example, in the superWIMP scenarios, one may probe gravitational interactions between fundamental particles and provide quantitative evidence for supergravity. In the WIMP scenarios, the thermal relic density studies simultaneously bound new contributions to dark energy at the time of freezeout, probing variations in the strength of gravity at $\\sim 1$ ns after the Big Bang, and possibly shedding light on the dark energy problem. It is rather striking that in many of these scenarios, the LHC, along with other observatories and experiments, could solve many old questions, such as the identity and origin of dark matter. If any of the ideas discussed here is realized in nature, the interplay of collider physics with cosmology and astrophysics in the next few years will likely yield profound insights about the Universe, its contents, and its evolution. \\ack I thank the organizers of GRG18/Amaldi7 for the invitation to participate in this stimulating conference and my collaborators for their many insights regarding the work discussed here. This work was supported in part by NSF Grants PHY--0239817 and PHY--0653656, NASA Grant NNG05GG44G, and the Alfred P.~Sloan Foundation." }, "0801/0801.4511_arXiv.txt": { "abstract": "{Blue Stragglers Stars (BSSs) are thought to form in globular clusters by two main formation channels: $i)$ mergers induced by stellar collisions and $ii)$ coalescence or mass-transfer between companions in binary systems. The detailed study of the BSS properties is therefore crucial for understanding the binary evolution mechanisms, and the complex interplay between dynamics and stellar evolution in dense stellar systems.} {We present the first comparison between the BSS specific frequency and the binary fraction in the core of a sample of Galactic globular clusters, with the aim of investigating the relative efficiency of the two proposed formation mechanisms.} {We derived the frequency of BSSs in the core of thirteen low-density Galactic globular clusters by using deep ACS@HST observations and investigated its correlation with the binary fraction and various other cluster parameters.} {We observed a correlation between the BSS specific frequency and the binary fraction. The significance of the correlation increases by including a further dependence on the cluster central velocity dispersion.} {We conclude that the unperturbed evolution of primordial binaries could be the dominant BSS formation process, at least in low-density environments.} ", "introduction": "Blue Straggler Stars (BSSs) are objects that, in the color-magnitude diagram (CMD) of evolved stellar populations, lie along an extension of the Main Sequence (MS), in a region which is brighter and bluer than the Turn-off (TO). First discovered by Sandage (1953) in M3, they have been observed in all Galactic globular clusters (GCs; Piotto et al. 2004), in the field population (Carney et al. 2005), and in dwarf galaxies of the local group (Momany et al. 2007). Their location in the CMD suggests that BSSs have masses of $1.2 \\div 1.5~M_{\\odot}$, significantly larger than those of normal stars in old stellar systems (like GCs). Thus, they are thought to have increased their mass during their evolution. Two mechanisms have been proposed for their formation: $i)$ the merger of two stars induced by stellar collision (COL-BSSs; Hills \\& Day 1976) and $ii)$ coalescence or mass-transfer between two companions in a binary system (MT-BSSs; McCrea 1964). The two formation channels are thought to act with different efficiencies according to the cluster structural parameters (Fusi Pecci et al. 1992) and they can work simultaneously within the same cluster in different radial regions, corresponding to widely different stellar densities (Ferraro et al. 1997; Mapelli et al. 2006). Indeed, collisions are more frequent in the central region of GCs, because of the high stellar density, while MT-BSSs mainly populate the cluster periphery, where binary systems can more easily evolve in isolation without suffering exchange or ionization due to gravitational encounters. The whole scenario is further complicated by the cluster dynamical evolution that leads massive systems (like binaries and BSSs) to sink toward the cluster center in a time-scale comparable to the cluster relaxation time. A possible tool for distinguishing COL-BSSs from MT-BSSs is based on high-resolution spectroscopic analysis. In fact, anomalous chemical abundances are expected at the surface of BSSs resulting from mass-transfer activity (Sarna \\& de Greve 1996), while they are not predicted for COL-BSSs (Lombardi et al. 1995). However, such studies have just become feasible and they are limited to only a small number of BSSs in just one cluster (47 Tucanae; Ferraro et al. 2006). As an alternative way for getting insights on the relative efficiency of the two formation mechanisms, here we investigate possible correlations between the BSS population and the host cluster properties. ", "conclusions": "We measured the BSS specific frequency in the core of thirteen low-density Galactic GCs and investigated its correlation with different dynamical and general cluster parameters. We found evidences that, at least in this density regime, binary-rich environments are more efficient in producing BSS. No correlations have been found with the cluster central density, concentration, stellar collision rate, and half-mass relaxation time, in agreement with the results of Piotto et al. (2004) and Leigh et al. (2007). These evidences indicate that the collisional channel for the BSS formation has a very small efficiency in low-density GCs, while the mechanisms involving the unperturbed evolution of binary systems are dominant. The higher significance of the trivariate correlation among the BSS frequency, the binary fractions and the cluster velocity dispersion, indicates that, for a given binary fraction, the BSS specific frequency decreases with increasing velocity dispersion. This finding might be connected with the effect of the cluster velocity dispersion in the dynamical evolution of binary systems. In fact, a small cluster velocity dispersion corresponds to a lower energy limit between soft and hard binaries\\footnote{A binary is defined {\\it soft} ({\\it hard}) if its binding energy ($E = -G~m_{1}~m_{2}/2~a$, with $a$ being the orbital separation of the two components) is smaller (larger) than the mean kinetic energy of normal cluster stars ($K = m~\\sigma_{\\rm v}^2$, with $m$ being the average mass of cluster stars).}, i.e., to a larger fraction of hard binary systems. Since the natural evolution of hard binaries is to increase their binding energy (i.e. decrease their orbital separation; Heggie 1975), this implies that low velocity dispersion GCs should host a larger fraction of hard (and close) binaries, able to both survive possible stellar encounters, and activate mass-transfer and/or merging processes between the companions. A larger fraction of BSSs formed by the evolution of primordial binaries is therefore expected in lower velocity dispersion GCs (see also Davies et al. 2004). Such an effect of $\\sigma_{\\rm v}$ (in terms of both hardening and shrinking the binary systems) might be, in turn, at the origin of the inverse correlation between the BSS frequency and the cluster total luminosity (mass) observed in open clusters (De Marchi et al. 2006), low density GCs (Sandquist 2005), as well as high density GCs (Piotto et al. 2004; Leigh et al. 2007). Indeed, the more massive GCs have larger central velocity dispersions (as a consequence of the virial theorem for systems with similar radii, as GCs; see Fig. 1 of Djorgovski 1995). For most of these clusters, however, the binary fraction is still unknown and the trivariate correlation between $F$, $\\xi_{bin}$ and $\\sigma_{\\rm v}$ cannot be derived. If its significance and its interpretation in terms of the velocity dispersion effect is confirmed also in high-density clusters, then stellar collisions might play a secondary role in the production of BSSs, and the evolution of primordial binaries should always be the dominant process. For the moment, however, these conclusions remain speculative, since the sample of GCs analyzed here, in spite of being the largest to date with known binary fraction, is still too small for statistically reliable assessments. Enlarging the sample of GCs with known binary and BSS fraction is therefore essential and urgent to verify the findings presented in this paper on a more robust statistical basis." }, "0801/0801.1116_arXiv.txt": { "abstract": "We describe new, deep MIPS photometry and new high signal-to-noise optical spectroscopy of the 2.5 Myr-old IC 348 Nebula. To probe the properties of the IC 348 disk population, we combine these data with previous optical/infrared photometry and spectroscopy to identify stars with gas accretion, to examine their mid-IR colors, and to model their spectral energy distributions. IC 348 contains many sources in different evolutionary states, including protostars and stars surrounded by primordial disks, two kinds of transitional disks, and debris disks. Most disks surrounding eary/intermediate spectral-type stars ($>$ 1.4 M$_{\\odot}$ at 2.5 Myr) are debris disks; most disks surrounding solar and subsolar-mass stars are primordial disks. At the 1--2 $\\sigma$ level, more massive stars also have a smaller frequency of gas accretion and smaller mid-IR luminosities than lower-mass stars. These trends are suggestive of a stellar mass-dependent evolution of disks, where most disks around high/intermediate-mass stars shed their primordial disks on rapid, 2.5 Myr timescales. The frequency of MIPS-detected transitional disks is $\\approx$ 15--35\\% for stars plausibly more massive than 0.5 M$_{\\odot}$. The relative frequency of transitional disks in IC 348 compared to that for 1 Myr-old Taurus and 5 Myr-old NGC 2362 is consistent with a transition timescale that is a significant fraction of the total primordial disk lifetime. ", "introduction": "Most $\\le$ 1 Myr-old stars are surrounded by optically-thick, accreting {\\it primordial} disks, which produce strong near-to-mid infrared (IR) emission \\citep[L$_{d}$/L$_{\\star}$ $\\ge$ 0.1;][]{Kh95}. The gas and dust in these disks comprise the building blocks of planets. As stars age, the dust grains grow, settle toward the midplane, and become incorporated into planetesimals. Circumstellar gas depletes by accretion onto the star \\citep{Ha98}. All of these processes reduce the amount and frequency of disk emission on timescales of $\\sim$ 3--7 Myr for most stars \\citep{He07a, Hi08b,Clp09}. By $\\sim$ 10 Myr, most stars do not have optically-thick primordial disks \\citep{Cu07a,Hi08b}. Disks around these older stars are typically debris disks \\citep{Cu08a}, which have weaker, optically-thin dust emission (L$_{d}$/L$_{\\star}$ $\\lesssim$ 10$^{-3}$) and lack evidence for circumstellar gas accretion. Because stellar radiation -- radiation pressure and Poynting-Robertson drag -- or stellar wind drag can remove dust in debris disks on timescales much less than the age of the star, their dust requires a replenishment source, which is supplied by active icy or rocky planet formation \\citep[e.g.][]{Bp93, Kb04, Kb08}. Because debris disks lack copious amounts of circumstellar gas needed to form gas giant planets, identifying when most primordial disks turn into debris disks pinpoints an empirical upper limit for the formation timescale for gas giant planets. Recent \\textit{Spitzer Space Telescope} studies of 1--5 Myr-old clusters indicate that primordial disks disappear faster around early type, high/intermediate-mass stars than they do around late type, low-mass stars \\citep{Ca06, La06, He07a, Clp09}. By 5 Myr, debris disks completely dominate the disk population around high/intermediate-mass stars. At 5 Myr, most disks around lower-mass stars appear to have inner holes and/or extremely low warm dust masses \\citep{Dahm09,Clp09}, which indicates that they may be disappearing and actively making the primordial-to-debris disk transition. Thus, the epoch at which most high/intermediate-mass stars make the primordial-to-debris disk transition must occur at an age prior to 5 Myr and may be identified by mid-IR observations of younger clusters. With a median age of $\\sim$ 2--3 Myr, stars in the IC 348 Nebula \\citep{He98} may help to constrain the epoch when most primordial disks around high/intermediate-mass stars turn into debris disks. Cluster membership is well determined \\citep[e.g.][]{He98,Lu98,Lu03,Mu07}. At only 320 pc distant, the cluster provides a sensitive probe of low-levels of disk emission in spite of its high infrared background. Recently, \\citet{La06} and \\citet{Muzerolle2006} analyzed Spitzer Cycle 1 IRAC and MIPS IC 348 data and found evidence for substantial disk evolution. Based on the IRAC 3.6 $\\mu m$ to 8 $\\mu m$ flux slope, \\citet{La06} show that only $\\approx$ 30\\% of cluster stars have strong ('thick') IRAC excess emission comparable to stars with primordial disks (e.g. most disk-bearing stars in Taurus). Another $\\approx$ 23\\% have weaker ('anemic') IRAC excess emission, and $\\approx$ 44\\% lack IRAC excess emission ('diskless'). However, inferring the physical state of disks from the observed IRAC slope is not straightforward. Both debris disks and 'evolved' primordial disks ('transitional' disks: those with lower disk masses and/or inner holes) can have weak IRAC-excess emission \\citep{Cu07b, Clp09}. Furthermore, while SED analysis including longer wavelength data (e.g. 24 $\\mu m$) helps to break this degeneracy \\citep{Clp09}, only $\\approx$ 30\\% of IC 348 stars have Cycle 1 MIPS detections. Therefore, deeper MIPS data coupled with other diagnostics of the disk evolutionary state (e.g. gas accretion signatures) would help to make IC 348 a better laboratory for studying disk evolution. In this paper, we combine new, high signal-to-noise optical spectra and new, deeper MIPS 24 $\\mu m$ and 70 $\\mu m$ photometry of IC 348 with archival data to study disk evolution in more detail. We first use the optical spectra to search for evidence of circumstellar gas accretion, which is absent in debris disks. Deeper MIPS data is then used to establish a much larger sample of disk-bearing stars. From these data, we compare the disks SEDs to models to constrain their evolutionary states, to compare disk properties around high-mass stars to low-mass stars, and to investigate the timescale for and the morphology of the primordial-to-debris disk transition. ", "conclusions": "\\subsection{Summary of Results} Using new optical spectra and new, deep MIPS 24 $\\mu m$ and 70 $\\mu m$ photometry we investigated the disk population of the 2.5 Myr-old IC 348 Nebula. Combining these data with previous work from \\citet{La06}, we analyzed optical spectroscopic data for all stars, performed SED modeling, and examined the mid-IR colors of MIPS-detected members to constrain the disks' evolutionary states and probe how disk properties vary with stellar mass. Our study yields the following major results: \\begin{itemize} \\item IC 348 sources with MIPS-70 $\\mu m$ detections include flat-spectrum protostars and pre-main sequence stars with optically-thick, luminous far IR disk emission. Some optically-thick disks show evidence for depleted inner disks; others do not. \\item IC 348 stars with new and previous MIPS detections have disks in evolutionary states ranging from primordial disks to debris disks. In agreement with recent work \\citep{La06, He07a, Clp09}, we find evidence for two separate pathways from primordial disks to debris disks. In homologously depleted disks, disks deplete their reservoir of small dust grains at all disk locations simultaneously. In a second sequence ('disks with inner holes'), disks clear their supply of small dust grains from the inside out. \\item At a $\\sim$ 1--2 $\\sigma$ significance, signatures of circumstellar gas accretion are more frequent for solar and subsolar-mass stars than for high/intermediate-mass stars. This result is consistent with a gas disk dispersal timescale that is shortest for high/intermediate-mass stars. \\item The mid-IR disk luminosities of MIPS-detected disks are stellar-mass dependent. Relative to the stellar photosphere, 24 $\\mu m$ emission from disks is lower for high/intermediate-mass stars than for solar/subsolar-mass stars. If disks generally decline in luminosity as a function of time, this result implies that disks around high/intermediate-mass stars evolve faster. \\item The evolutionary states of MIPS-detected disks are also stellar mass dependent. Most MIPS-detected disks surrounding high/intermediate-mass stars stars appear to be debris disks; primordial disks comprise only $\\sim$ 15\\% of the disk population for M$_{\\star}$ $>$ 1.4 M$_{\\odot}$. In contrast, most MIPS-detected disks around solar and subsolar-mass stars are primordial disks. For stars of all masses, transitional disks (homologously depleted or disks with inner holes) comprise $\\sim$ 15--35\\% of the MIPS-detected disk population. \\end{itemize} \\subsection{The Evolutionary State of \\textit{Anemic} Disks} Our SED modeling and investigation of the mid-IR colors of IC 348 stars clarifies the nature of \\textit{anemic} disks identified by \\citet{La06}. Among late type stars, most anemic disks are homologously depleted disks or are disks with inner holes. Several anemic disks are probably primordial disks. Among earlier-type stars with probable masses $\\gtrsim$ 1.4 M$_{\\odot}$, anemic disks comprise a broader range of evolutionary states. At least two anemic disks (surrounding IDs 6 and 8) are probably debris disks. In contrast, ID-31 has strong mid-IR emission and gas accretion signatures more consistent with a transitional disk, specifically one with an inner hole. Thus, as pointed out by \\citet{Clp09}, a single IRAC slope serves as a useful first-order probe of disk properties, but a full analysis of IRAC \\textit{and} MIPS data combined with gas accretion signatures is required for an accurate taxonomy of disks. Our results indicate that a full analysis is especially necessary if the stellar population includes both high/intermediate-mass stars and lower-mass stars. Given the diversity in disk properties for 2.5 Myr-old IC 348, we suggest that a single flux slope may be best suited as a probe of disk evolution in the youngest regions (e.g. Taurus; NGC 1333), where less diversity in disk states (e.g. few if any debris disks) is expected. \\subsection{Constraints on the Primordial-to-Debris Disk Transition} Recent Spitzer studies have placed strong constraints on the timescales for the evolution of primordial disks into debris disks. By 5 Myr, most high/intermediate-mass stars and solar-mass stars either lack evidence for a disk or have disk properties suggestive of a debris disk \\citep{Ca06, He08, Clp09}. At this age, many (detectable) subsolar-mass stars have weaker levels of mid-IR disk emission than typical primordial disks \\citep{Clp09,Dahm09}. This work shows that by 2.5 Myr most disks around high/intermediate-mass stars are either actively leaving the primordial disk phase or have already reached the debris disk phase. Together with previous results \\citep[e.g.][]{Ca06, He07a, Clp09}, our results clearly support a stellar-mass dependent timescale for the disappearance of primordial disks and the emergence of debris disks. Within the context of planet formation, gas giant planets around high/intermediate-mass stars have much less time to form than around low-mass stars. If gas giant planets are more frequent around high/intermediate-mass stars \\citep{Jj07}, their formation must be very rapid and efficient. Only recently have realistic models \\citep{Kb09} that form gas giant planets via core accretion been successful in forming the cores of such planets in the short timescales implied from this paper (2.5 Myr). Some trends in exoplanet properties, such as the semimajor axis distribution, may emerge from the competing effects of core formation timescales and primordial disk dispersal timescales \\citep{Currie2009}. The fraction of IC 348 disks in a transitional phase serves as a useful contrast to results obtained for younger clusters like Taurus and older clusters like NGC 2362. Based primarily on IRAS data, the computed frequency of transitional disks in Taurus is small, $<<$ 10\\% \\citep[e.g.][]{Sk90, Sp95, Ww96}. These authors then argued that the lifetime of transitional disks is $\\approx$ 1--10\\% the age of Taurus: $\\approx$ 0.01--0.1 Myr. If this inference were correct, the total disk lifetime may be several Myr, but most of the disk is dissipated rapidly. However, analysis of Spitzer data for 5 Myr-old NGC 2362 finds a much higher frequency of transitional disks \\citep{Clp09}. Both types of transitional disks (inner holes, homologously depleted) greatly outnumber primordial disks. This result argues for far longer typical transition timescale ($\\approx$ 1 Myr). In the intermediate age IC 348, the frequency of transition disks is intermediate between the frequency for 1 Myr-old Taurus and 5 Myr-old NGC 2362 for solar/subsolar-mass stars. Thus, the number of transition disks relative to primordial disks around solar/subsolar-mass stars appears to be an increasing function of stellar age. This result is expected if the typical transition timescale is an appreciable fraction of the typical primordial disk lifetime. Thus, the transition disks in Taurus and IC 348 could remain in such a state for an extended period. However, it is possible that transition disk lifetimes, like primordial disk lifetimes, also have an intrinsic dispersion (i.e., a gaussian distribution of lifetimes), where sources in Taurus represent those that make the primordial-to-debris disk transition fastest and sources in 5--10 Myr-old clusters make the transition the slowest. Quantifying the range of times a disk spends in a transitional phase and whether the typical timescale depends on stellar mass requires observations of many more 1--10 Myr-old clusters. While debris disks typically only have excess emission longwards of $\\approx$ 20 $\\mu m$, the debris disk population in IC 348 contains at least two sources (ID-6 and ID-8) that have warmer dust more indicative of terrestrial planet formation than icy planet formation. Thus, warm debris disks consistent with the observable signatures of terrestrial planet formation may emerge as early as $\\approx$ 2.5 Myr around high/intermediate-mass stars. Observations of more 2--5 Myr-old clusters are needed to determine if the frequency of warm debris disks is higher than the $\\sim$ 4\\% derived for 10--15 Myr-old clusters \\citep{Cu07a, Cu08}. The existence of both warm debris disks and colder debris disks (lacking IRAC excess emission) suggests that even at 2.5 Myr debris disk populations may exhibit a range of dust temperatures consistent with a range of locations over which the debris-producing stages of planet formation are ongoing." }, "0801/0801.4505_arXiv.txt": { "abstract": "We present measurements with the VLBA of the variability in the centroid position of \\SgrA\\ relative to a background quasar at $7\\,\\mm$ wavelength. We find an average centroid wander of $71\\pm 45\\,\\muas$ for time scales between 50 and $100\\,\\min$ and $113\\pm50\\,\\muas$ for timescales between 100 and $200\\,\\min$, with no secular trend. These are sufficient to begin constraining the viability of the accretion hot-spot model for the radio variability of \\SgrA. It is possible to rule out hot spots with orbital radii above $15\\,G M_{\\rm Sgr A*}/c^2$ that contribute more than 30\\% of the total $7\\,\\mm$ flux. However, closer or less luminous hot spots remain unconstrained. Since the fractional variability of \\SgrA\\ during our observations was $\\sim20$\\% on time scales of hours, the hot-spot model for \\SgrA's radio variability remains consistent with these limits. Improved monitoring of \\SgrA's centroid position has the potential to place significant constraints upon the existence and morphology of inhomogeneities in a supermassive black hole accretion flow. ", "introduction": "There is now overwhelming evidence that \\SgrA\\ is a supermassive black hole at the center of the Milky Way. Many stars are observed to orbit about a common focal position, requiring an unseen mass of $\\approx4\\times10^6$~\\Msun\\ contained within a radius of less than 100~AU \\citep{Schoedel:2002,Ghez:2003}, for a distance to the center of 8.0~kpc \\citep{Reid:1993}. Accurate registration of the infrared and radio reference frames \\citep{Menten:1997,Reid:2003} reveal that the common orbital focal position is coincident with \\SgrA\\ to within measurement uncertainty of $\\approx10$~mas. Finally, the absence of intrinsic motion of \\SgrA\\ at levels near that expected for a $4\\times10^6$~\\Msun\\ object \\citep{Reid-Brun:2004}, coupled with a size less than ~1~AU \\citep{Bower:2004}, provide a lower limit on mass density of $\\sim10^{22}$~\\Msun~pc$^{-3}$, which is only two orders of magnitude less than the density of a $4\\times10^6$~\\Msun\\ non-rotating black hole within its innermost stable orbit. There can now be little doubt that \\SgrA\\ is a supermassive black hole. \\citet{Reid-Brun:2004} present measurements of the position of \\SgrA\\ relative to a compact extragalactic radio source (\\EGS, also refered to as J1745-283 in earlier publications). These measurements were conducted with the NRAO \\footnote{The National Radio Astronomy Observatory is operated by Associated Universities, Inc., under a cooperative agreement with the National Science Foundation.} Very Long Baseline Array (VLBA) over a period of $8\\,\\yr$ at a wavelength of $7\\,\\mm$ ($43\\,\\GHz$) and have been used to determine the apparent proper motion of \\SgrA. Over time scales of months or longer, \\SgrA's {\\it apparent} motion is dominated by the effects of the orbit of the Sun about the center of the Galaxy. The component of the Sun's orbit in the Galactic plane is uncertain at roughly the 10\\% level, and this limits estimation of any intrinsic motion of \\SgrA\\ at about the $\\pm20$~\\kms\\ level. However, the component of the motion of the Sun out of the Galactic plane is known to high accuracy [$7.16\\pm0.38$~\\kms\\ toward the North Galactic Pole; \\citet{Dehnen-Binney:1998}]. After removing the effects of the Sun's motion, the residual motion of \\SgrA\\ perpendicular to the Galactic plane is very small, $\\lesssim 1\\,\\km\\,\\s^{-1}$, as expected for a supermassive black hole (SMBH) at the dynamical center of the Galaxy. While previous work concentrated on the long-term motion of \\SgrA, here we analyze its short-term position ``wander'' on time scales of hours to weeks. Short-timescale motion of the centroid position of \\SgrA\\ would be expected if a portion of the emission comes from material orbiting about the SMBH. The degree of centroid variability would necessarily depend upon the brightness of the orbiting material, the degree to which its emission is nonuniform, and the orbital radius dominating the total flux. We use a simple hot-spot model to relate the constraint from the observed short-term position wander of \\SgrA\\ (\\S \\ref{OoCM}) to a constraint upon the presence of strong inhomogeneities in the accretion flow onto the SMBH as a function of hot-spot luminosity and orbital period (\\S \\ref{Constraints}). Finally, concluding remarks are contained in \\S \\ref{C}. ", "conclusions": "\\label{C} Possible reasons for position wander include intrinsic variations in the position of the emitting plasma (\\eg, variations in the accretion flow or perhaps in a jet) or extrinsic processes such as refractive interstellar scattering. \\SgrA\\ is observed to be diffractively scattered to a size of $\\theta_{sc} \\sim 0.5 (\\lambda/0.7~{\\rm cm})^2$~mas, where $\\lambda$ is the observing wavelength. Flux density fluctuations are modest and decrease in strength with increasing wavelength; thus strong refractive scintillations are not indicated \\citep{Gwinn:1991}. Any refractive position wander should be $\\ll \\theta_{sc}$ and should occur on time scales $>\\theta_{sc}D/v$, where $D$ is the distance and $v$ is the transverse velocity of the scattering ``screen'' relative to the observer \\citep{Romani:1986}. For $D\\approx\\Ro\\approx8$~kpc \\citep{Reid:1993} and $v\\sim100$~\\kms, characteristic of material in the inner $\\sim100$~pc of the Galaxy where large scattering sizes are observed, the refractive time scale is $>10^3$~hours. Thus, we would not expect a significant contribution to the short-term wander of \\SgrA\\ from refractive scattering. For comparison, \\citet{Gwinn:1988}, using VLBI observations of the Sgr~B2(N) \\hho\\ masers near the Galactic center, find a wander limit of $<18~\\muas$ over timescales of months for maser spots, which are diffractively scattered to a comparable size (at 22 GHz) as \\SgrA\\ (at 43 GHz). Of course, our results provide an observation limit to any refractive position wander. Since extrinsic sources of position wander (scattering) are unlikely to be dominant, we now discuss the implications for intrinsic wander from variations in brightness within an accretion disk given in \\S\\ref{Constraints}. Our observations of the lack of short-term wander of the centroid position of \\SgrA\\ presented in \\S\\ref{section:hours} give an upper limit of $\\approx100~\\muas$ for time scales of $\\approx1~{\\rm to}~4$ hours. This translates to an upper limit on the wander versus orbital period plots in Fig.~\\ref{fig:drs} as indicated by the horizontal red line and hatched region. (In the very unlikely event that the accretion disk inclination is both near $90^\\circ$ and oriented nearly North-South on the sky, we would need to use our NS limits, which are a factor of three weaker.) Our EW limit is below the dotted blue line in Fig.~\\ref{fig:drs}, which is for hot spot flux density dominating over disk (or possible jet) emission, for orbital periods exceeding 120~min (corresponding to orbital radii larger than $15\\,G M_{\\rm Sgr A*}/c^2$ for $M_{\\rm Sgr A*}= 4\\times10^6$ \\Msun). For cases in which the hot spot flux density is weaker than that of the disk, somewhat longer periods are allowed. For example, for $F_{\\rm spot}/F_{\\rm disk} = 0.37$, orbital periods longer than 5~hr are excluded. In practice, the limits placed by current $7\\,\\mm$ VLBI are significantly weaker. The limited sensitivity to hot spots on compact orbits is primarily due to two reasons: (1) ``long'' integration times ($T\\gtrsim 1\\,\\hr$) average much of the short time variability out, and (2) the opacity of the accretion flow itself makes it difficult to view hot-spots on compact orbits at $7\\,\\mm$. The integration time is limited by the sensitivity issues and the small number of antennas yielding interferometer baselines $<1500$~km afforded by the current VLBA; higher bandwidth recording in the future should help alleviate this problem. The optical depth is a property of Sgr A* itself, and can only be addressed by observations at shorter wavelengths. However, even in the absence of an optically thick accretion flow, it is not possible to increase the centroid variability by more than an order of magnitude due to the intrinsically small orbital radii, as seen by comparing the blue and green limits in the lower-left panel of Fig. \\ref{fig:drs}. Nevertheless, high-resolution astrometry is reaching sensitivities and resolutions sufficient to begin to test the hot-spot model for bright Sgr A* flares. Unfortunately, the typical fractional variability at $7\\,\\mm$ during our observations was roughly $\\pm20\\%$, implying that significant improvement in positional accuracy will be required to constrain such events. Since the observed centroid wander is consistent with systematic errors, owing predominantly to centimeter-scale errors in the modeling of the atmospheric path-delays, substantially increasing accuracy will require better calibration techniques. However, for the somewhat rare instances in which the spot is substantially brighter \\citep{Zhao:2001}, the VLBA at 7, or possibly 3,~mm wavelength appears poised to provide significant limits upon the existence and morphology of inhomogeneities in the accretion flow surrounding \\SgrA. Ultimately, observations at $\\sim1$~mm wavelength with VLBI techniques or at infrared wavelengths with an instrument like GRAVITY \\citep{GRAVITY} may be necessary to image the region within $\\sim3$ Schwarzschild radii on the short time scales needed to test the hot-spot model. \\vskip 1truecm A.B. is supported by the Priority Programme 1177 of the Deutsche Forschungsgemeinschaft. \\vskip 0.5truecm {\\it Facilities:} \\facility{VLBA}" }, "0801/0801.1320_arXiv.txt": { "abstract": "{}{ To determine alpha effect and turbulent magnetic diffusivity for mean magnetic fields with profiles of different length scale from simulations of isotropic turbulence, and to relate these results to nonlocal formulations in which alpha and the turbulent magnetic diffusivity correspond to integral kernels. }{ A set of evolution equations for magnetic fields is solved which gives the response to imposed test fields, that is, mean magnetic fields with various wavenumbers. Both an imposed fully helical steady flow consisting of a pattern of screw-like motions (Roberts flow) and time-dependent statistically steady isotropic turbulence are considered. In the latter case the aforementioned evolution equations are solved simultaneously with the momentum and continuity equations. The corresponding results for the electromotive force are used to calculate alpha and magnetic diffusivity tensors. }{ For both the Roberts flow under the second--order correlation approximation and isotropic turbulence alpha and turbulent magnetic diffusivity are largest at large scales and these values diminish toward smaller scales. In both cases the alpha effect and turbulent diffusion kernels are well approximated by exponentials, corresponding to Lorentzian profiles in Fourier space. For isotropic turbulence the turbulent diffusion kernel is half as wide as the alpha effect kernel. For the Roberts flow beyond the second--order correlation approximation the turbulent diffusion kernel becomes negative at large scales. }{} ", "introduction": "\\label{Intro} Stars and galaxies harbour magnetic fields whose scales are large compared with the scale of the underlying turbulence. This phenomenon is successfully explained in terms of mean--field dynamo theory discussed in detail in a number of text books and reviews (e.g.\\ Moffatt 1978, Krause \\& R\\\"adler 1980, Brandenburg \\& Subramanian 2005a). In this context velocity and magnetic fields are split into large--scale and small--scale components, $\\UU=\\meanUU+\\uu$ and $\\BB=\\meanBB+\\bb$, respectively. The crucial quantity of the theory is the mean electromotive force caused by the small--scale fields, $\\meanEMF=\\overline{\\uu\\times\\bb}$. In many representations it is discussed under strongly simplifying assumptions. Often the relationship between the mean electromotive force and the mean magnetic field is tacitly taken as (almost) local and as instantaneous, that is, $\\meanEMF$ in a given point in space and time is considered as determined by $\\meanBB$ and its first spatial derivatives in this point only, and the possibility of a small--scale dynamo is ignored. Then the mean electromotive force is given by \\EQ \\meanemf_i=\\alpha_{ij}\\meanB_j+\\eta_{ijk} \\partial \\meanB_j / \\partial x_k \\label{meanemfi} \\EN with two tensors $\\alpha_{ij}$ and $\\eta_{ijk}$. If the turbulence is isotropic the two tensors are isotropic, too, that is $\\alpha_{ij}=\\alpha \\delta_{ij}$ and $\\eta_{ijk}=\\eta_{\\rm t} \\epsilon_{ijk}$ with two scalar coefficients $\\alpha$ and $\\eta_{\\rm t}$. Then the expression \\eq{meanemfi} simplifies to \\EQ \\meanEMF=\\alpha\\meanBB-\\eta_{\\rm t}\\meanJJ \\, , \\label{meanEMF} \\EN where we have denoted $\\nab \\times \\BB$ simply by $\\JJ$ (so that $\\JJ$ is $\\mu_0$ times the electric current density, where $\\mu_0$ is the magnetic permeability of free space). The coefficient $\\alpha$ is, unlike $\\eta_{\\rm t}$, only non--zero if the turbulence lacks mirror--symmetry. The coefficient $\\eta_{\\rm t}$ is referred to as the turbulent magnetic diffusivity. In general, the mean electromotive force has the form \\EQ \\meanEMF=\\meanEMF_0 + \\KK \\circ \\meanBB, \\label{kernel} \\EN where $\\meanEMF_0$ stands for a part of $\\meanEMF$ that is independent of $\\meanBB$, and $\\KK \\circ \\meanBB$ denotes a convolution in space and time of a kernel $\\KK$ with $\\meanBB$ (see, e.g., Krause \\& R\\\"adler 1980, R\\\"adler 2000, R\\\"adler \\& Rheinhardt 2007). Due to this convolution, $\\meanEMF$ in a given point in space and time depends on $\\meanBB$ in a certain neighborhood of this point, with the exception of future times. This corresponds to a modification of \\eq{meanemfi} such that also higher spatial and also time derivatives of $\\meanBB$ occur. In this paper we ignore the possibility of coherent effects resulting from small--scale dynamo action and put therefore $\\meanEMF_0$ equal to zero. For the sake of simplicity we assume further the connection between $\\meanEMF$ and $\\meanBB$ to be instantaneous so that the convolution $\\KK \\circ \\meanBB$ refers to space coordinates only. The memory effect, which we so ignore, has been studied previously by solving an evolution equation for $\\meanEMF$ (Blackman \\& Field 2002). For homogeneous isotropic turbulence we may then write, analogously to (\\ref{meanEMF}), \\EQ \\meanEMF=\\hat{\\alpha}\\circ\\meanBB-\\hat{\\eta}_{\\rm t}\\circ\\meanJJ, \\label{meanEMFzz} \\EN or, in more explicit form, \\EQ \\meanEMF(\\xx)=\\int\\left[\\hat{\\alpha}(\\xi)\\meanBB(\\xx-\\xxi) -\\hat{\\eta}_{\\rm t}(\\xi)\\meanJJ(\\xx-\\xxi)\\right]\\,\\dd^3 \\xi \\label{meanEMFzzprime} \\EN with two functions $\\hat{\\alpha}$ and $\\hat{\\eta}_{\\rm t}$ of $\\xi = |\\xxi|$ that vanish for large $\\xi$. The integration is in general over all $\\xi$--space. Although $\\meanEMF$ and $\\meanBB$ as well as $\\hat{\\alpha}$ and $\\hat{\\eta}_{\\rm t}$ may depend on time the argument $t$ is dropped everywhere. For a detailed justification of the relations \\eq{meanEMFzz} and \\eq{meanEMFzzprime} we refer to \\App{justific}. In the limit of a weak dependence of $\\meanBB$ and $\\meanJJ$ on space coordinates, i.e.\\ when the variations of $\\meanBB (\\xx - \\xxi)$ and $\\meanJJ (\\xx - \\xxi)$ with $\\xxi$ are small in the range of $\\xi$ where $\\hat{\\alpha} (\\xi)$ and $\\hat{\\eta}_{\\rm t} (\\xi)$ are markedly different from zero, the relations \\eq{meanEMFzz} or \\eq{meanEMFzzprime} turn into \\eq{meanEMF}, and we see that \\mbox{$\\alpha = \\int \\hat{\\alpha} (\\xi) \\, \\dd^3 \\xi$} and $\\eta_{\\rm t} = \\int \\hat{\\eta}_{\\rm t} (\\xi) \\, \\dd^3 \\xi$. At first glance the representations \\eq{meanEMFzz} and \\eq{meanEMFzzprime} of $\\meanEMF$ look rather different from \\eq{kernel}. Considering $\\meanJJ = \\nab \\times \\meanBB$ and carrying out an integration by parts we may however easily rewrite \\eq{meanEMFzzprime} into \\EQ \\meanemf_i (\\xx) = \\int K_{ij} (\\xxi) \\meanB_j (\\xx - \\xxi) \\, \\dd^3\\xi \\label{eq1} \\EN with \\EQ K_{ij} (\\xxi) = \\hat{\\alpha} (\\xi) \\delta_{ij} - \\frac{1}{\\xi} \\frac{\\partial \\hat{\\eta}_{\\rm t} (\\xi)}{\\partial \\xi} \\epsilon_{ijk} \\xi_k \\, . \\label{eq3} \\EN We further note that due to the symmetry of $\\hat{\\alpha} (\\xi)$ in $\\xxi$ only the part of $\\meanBB (\\xx - \\xxi)$ that is symmetric in $\\xxi$, i.e.\\ the part that can be described by $\\meanBB (\\xx)$ and its derivatives of even order, contributes to the $\\hat{\\alpha}$ terms in \\eq{meanEMFzzprime} or in \\eq{eq1} and \\eq{eq3}. The symmetry of $\\hat{\\eta}_{\\rm t} (\\xi)$ implies that only the part of $\\meanBB (\\xx - \\xxi)$ antisymmetric in $\\xxi$, which corresponds to the derivatives of $\\meanBB (\\xx)$ of odd order, contributes to the $\\hat{\\eta}_{\\rm t}$ terms. Finally, referring to a Cartesian coordinate system $(x,y,z)$ we define mean fields by averaging over all $x$ and $y$ so that in particular $\\meanEMF$ and $\\meanBB$ depend only on $z$ and on time. Then \\eq{meanEMFzzprime} turns into \\EQ \\meanEMF(z) =\\int\\left[\\hat{\\alpha}(\\zeta)\\meanBB(z-\\zeta) -\\hat{\\eta}_{\\rm t}(\\zeta)\\meanJJ(z-\\zeta)\\right]\\,\\dd \\zeta \\, . \\label{meanEMFzzprime2} \\EN The functions $\\hat{\\alpha} (\\zeta)$ and $\\hat{\\eta}_{\\rm t} (\\zeta)$ are just averages of $\\hat{\\alpha} (\\xi)$ and $\\hat{\\eta}_{\\rm t} (\\xi)$ over all $\\xi_x$ and $\\xi_y$. They are therefore real and symmetric in $\\xi_z\\equiv\\zeta $. The integration in \\eq{meanEMFzzprime2} is in general over all $\\zeta$. The remark on the limit of weak dependences of $\\meanBB$ and $\\meanJJ$ on space coordinates made above in connection with \\eq{meanEMFzz} and \\eq{meanEMFzzprime} applies analogously to \\eq{meanEMFzzprime2}. We have now $\\alpha = \\int \\hat{\\alpha} (\\zeta) \\, \\dd \\zeta$ and $\\eta_{\\rm t} = \\int \\hat{\\eta}_{\\rm t} (\\zeta) \\, \\dd \\zeta$. Relation \\eq{meanEMFzzprime2} can also be brought in a form analogous to \\eq{eq1} and \\eq{eq3}, \\EQ \\meanemf_i (z) = \\int K_{ij} (\\zeta) \\meanB_j (z - \\zeta) \\, \\dd \\zeta \\label{eq5} \\EN with \\EQ K_{ij} (\\zeta) = \\hat{\\alpha} (\\zeta) \\delta_{ij} - \\frac{\\partial \\hat{\\eta}_{\\rm t} (\\zeta)}{\\partial \\zeta} \\epsilon_{ij3} \\, . \\label{eq7} \\EN The remarks made under \\eq{eq1} and \\eq{eq3} apply, now due to the symmetries of $\\hat{\\alpha} (\\zeta)$ and $\\hat{\\eta}_{\\rm t} (\\zeta)$ in $\\zeta$, analogously to \\eq{meanEMFzzprime2}, \\eq{eq5} and \\eq{eq7}. It is useful to consider in addition to \\eq{meanEMFzzprime2} also the corresponding Fourier representation. We define the Fourier transformation in this paper by $Q(z) = \\int \\tilde{Q}(k) \\exp(\\ii k z) \\, \\dd (k/2 \\pi)$. Then this representation reads \\EQ \\tilde{\\meanEMF}(k) = \\tilde{\\alpha}(k) \\tilde{\\meanBB}(k) -\\tilde{\\eta}_{\\rm t}(k) \\tilde{\\meanJJ}(k) \\, . \\label{meanEMFzzprime3} \\EN Both $\\tilde{\\alpha}(k)$ and $\\tilde{\\eta}_{\\rm t}(k)$ are real quantities, and they are symmetric in $k$. The limit of weak dependences of $\\meanBB$ and $\\meanJJ$ on $z$ corresponds here to $k \\to 0$, and we have $\\alpha = \\tilde{\\alpha} (0)$ and $\\eta_{\\rm t} = \\tilde{\\eta}_{\\rm t} (0)$. Detailed analytic expressions for $\\hat{\\alpha}(\\zeta)$ and $\\hat{\\eta}_{\\rm t}(\\zeta)$, or $\\tilde{\\alpha}(k)$ and $\\tilde{\\eta}_{\\rm t}(k)$, can be derived, e.g., from results presented in Krause \\& R\\\"adler (1980). A numerical determination of quantities corresponding to $\\hat{\\alpha} (\\zeta)$ and $\\hat{\\eta}_{\\rm t} (\\zeta)$ has been attempted by Brandenburg \\& Sokoloff (2002) for shear flow turbulence. In this paper two specifications of the velocity field $\\uu$ will be considered. In the first case $\\uu$ is chosen such that it corresponds to a steady Roberts flow, which is periodic in $x$ and $y$ and independent of $z$. A mean--field theory of a magnetic field in fluid flows of this type, that are of course different from genuine turbulence, has been developed in the context of the Karlsruhe dynamo experiment (R\\\"adler et al.\\ 2002a,b, R\\\"adler \\& Brandenburg 2003). It turned out that the mean electromotive force $\\meanEMF$, except its $z$ component, satisfies relation \\eq{meanEMF} if any nonlocality in the above sense is ignored (see also \\App{robdyn}). Several analytical and numerical results are available for comparison with those of this paper. In the second case $\\uu$ is understood as homogeneous, isotropic, statistically steady turbulence, for which the above explanations apply immediately. Employing the method developed by Schrinner et al.\\ (2005, 2007) we will in both cases numerically calculate the functions $\\tilde{\\alpha}(k)$ and $\\tilde{\\eta}_{\\rm t}(k)$ as well as $\\hat{\\alpha}(\\zeta)$ and $\\hat{\\eta}_{\\rm t}(\\zeta)$. ", "conclusions": "The test field procedure turned out to be a robust method for determining turbulent transport coefficients (see Brandenburg 2005, Sur et al.\\ 2008 and Brandenburg et al.\\ 2008). The present paper shows that this also applies to the Fourier transforms of the integral kernels which occur in the nonlocal connection between mean electromotive force and mean magnetic field, in other words to the more general scale--dependent version of those transport coefficients. For isotropic turbulence the kernels $\\hat\\alpha$ and $\\hat\\eta_{\\rm t}$ have a dominant large-scale part and decline monotonously with increasing wavenumbers. This is consistent with earlier findings (cf.\\ Brandenburg \\& Sokoloff 2002), where however the functional form of the decline remained rather uncertain. Our present results suggest exponential kernels, corresponding to Lorentzian profiles in wavenumber space. The kernel for the turbulent magnetic diffusivity is about half as wide as that for the alpha effect. This result is somewhat unexpected and would be worthwhile to confirm before applying it to more refined mean field models. On the other hand, the effects of nonlocality become really important only when the scale of the magnetic field variations is comparable or smaller than the outer scale of the turbulence. One of the areas where future research of nonlocal turbulent transport coefficients is warranted is thermal convection. Here the vertical length scale of the turbulent plumes is often comparable to the vertical extent of the domain. Earlier studies by Miesch et al.\\ (2000) on turbulent thermal convection confirmed that the transport of passive scalars is nonlocal, but it is also more advective than diffusive. It may therefore be important to also allow for nonlocality in time. This would make the expansion of passive scalar perturbations more wave-like, as was show by Brandenburg et al.\\ (2004) using forced turbulence simulations." }, "0801/0801.4675_arXiv.txt": { "abstract": "{} {We present a 0.2--200 keV broad-band study of absorbed AGN observed with \\textit{INTEGRAL}, \\textit{XMM-Newton}, \\textit{Chandra} and \\textit{ASCA} to investigate the continuum shape and the absorbing/reflecting medium properties.} {The sources are selected in the \\integral\\ AGN sample to have a 20--100 keV flux below 8$\\times$10$^{-11}$ $\\flux$ (5 mCrab), and are characterized by a 2--10 keV flux in the range (0.8--10)$\\times$10$^{-11}$ $\\flux$. The good statistics allow us a detailed study of the intrinsic and reflected continuum components. In particular, the analysis performed on the combined broad-band spectra allow us to investigate the presence of Compton reflection features and high energy cut-off in these objects.} {The column density of the absorbing gas establishes the Compton thin nature for three sources in which a measure of the absorption was still missing. The Compton thin nature of all the sources in this small sample is also confirmed by the diagnostic ratios F$_{\\rm x}$/F$\\left[\\rm O III\\right] $. The Compton reflection components we measure, reflection continuum and iron line, are not immediately compatible with a scenario in which the absorbing and reflecting media are one and the same, i.e. the obscuring torus. A possible solution is that the absorption is more effective than reflection, e.g. under the hypothesis that the absorbing/reflecting medium is not uniform, like a clumpy torus, or that the source is observed through a torus with a very shallow opening angle. The high energy cut-off (a lower limit in two cases) is found in all sources of our sample and the range of values is in good agreement with that found in type 1 Seyfert galaxies. At lower energies there is clear evidence of a soft component (reproduced with a thermal and/or scattering model), in six objects.} {} ", "introduction": "The broad-band properties of Seyfert 2 AGN have been extensively studied in the past years mainly using \\sax\\ data (\\cite{risaliti02}). The continuum shape in these objects can be reproduced by a steep power-law with photon index $\\Gamma \\sim$ 2. This model matches well a Comptonization scenario in which seed photons from a cold gas (the accretion disc) are up-scattered in a hot corona of relativistic electrons (\\cite{HM91}, \\cite{HMG97}). A photoelectric cut-off is present at X-ray energies and its value depends on the absorbing column density of the source. For N$_H$ in the $10^{22}-10^{23}$ \\cmM2 range the cut-off is below 2 keV, for N$_H$ in the $10^{23}-10^{24}$ \\cmM2 range the cut-off is in 3--10 keV band and for N$_H >\\sigma_{T}^{-1}=1.5\\times 10^{24}$ \\cmM2 (i.e. the Compton thick regime with $\\sigma_{T}^{-1}$ being the inverse of the Thomson cross section), the continuum can be observed only through its reflected component above 10 keV. The presence of a high energy cut-off is still an open issue. Risaliti (2002) found that the high energy cut-off is present only in $\\sim$ 30 per cent of a sample of 20 bright Sy 2 observed with \\sax. Above 10 keV the presence of a Compton reflection component has often been observed, however is its origin still unclear as well as if the reflecting/absorbing medium are one and the same (the putative ''molecular torus''). A cold iron line at 6.4 keV is often associated to this continuum Compton reflection hump. The IBIS gamma-ray imager on board \\integral\\ is surveying the whole sky above 20 keV with a mCrab ($\\sim $10$^{-11}$ erg cm$^{-2}$ s$^{-1}$) sensitivity in well exposed regions. With its angular resolution of 12 arcmin and the point source location accuracy of 1-2 arcmin for moderately bright sources (Ubertini et al. 2003) \\integral\\ has already detected a number of AGN, a large fraction of which were previously unknown as high energy emitters (\\cite{bird06}, \\cite{bassani06a}, \\cite{bird07}). In this paper we present a deep broad-band spectral analysis of seven absorbed AGN detected by \\integral. The sample is extracted from the Bassani et. al (2006a) survey, updated to include a number of optical classifications obtained afterwards (see \\cite{bassani06b}). The sample includes Seyfert 2 galaxies having a 20--100 keV flux below 5 mCrab and with X-ray data available at the time that this work started; from the sample we have excluded sources already studied by \\sax\\ except for one object (\\eso) which was retained in order to provide a direct comparison between this work and previous studies using \\sax\\ data and therefore an \\textit{a posteriori} check of our analysis procedure which is affected by the limitation of using non simultaneous X and soft gamma-rays data. Overall we can conclude that the sample used in this study is representative of the population of type 2 AGN detected by \\integral\\ above 10 keV; further broad band X/gamma-ray analysis on a complete sample of type 2 AGN selected in the 20-40 keV band is on going and will be the objective of a future paper\\footnote{6 other type 2 AGN from Bassani et al. survey are below the assumed threshold flux: 3 (IGR J12415+5750, IGR J20286+2544 and IGR J17513-2011) had no X-ray available data when this analysis started while 3 (NGC 1068, Mkn 3 and NGC 6300) had BeppoSAX observational results published but all are peculiar objects the first two being Compton thick AGN and the third one the prototype of \"changing look\" Seyfert 2 ( i.e. a source that goes from thick to thin and viceversa).}. Four sources (IGR J12391-1610=\\leda\\ hereafter, \\075, \\120 and \\j104) have been discovered first in soft gamma-ray band and immediately after observed with \\chandra\\ (\\leda, \\075 and \\120) and \\xmm\\ (\\j104). Two objects (\\eso\\ and \\ic) are known AGN (although \\ic\\ was never reported at X-ray energies before) and follow up observations with \\xmm\\ have been used in our broad-band analysis. For one object (\\ngc) archival \\asca\\ data have been used in this work. Using X-ray measurements in conjunction with \\integral\\ data allow us to build a spectrum in three decades of energy, and to distinguish between the various spectral components characterizing the broad-band spectra of absorbed AGN. This in turn will allow us, in particular, to study the absorption/reflection medium properties of the sample. These good quality spectra also permit a detailed investigation of the intrinsic continuum slope including the cut-off if present. The broad--band coverage available is a powerful and unique tool to address all these issues, making possible a simultaneous measurement of the intrinsic continuum (both photon index and high energy cut-off) and of the reflected continuum together with the Fe line. Note that a number of sources in the sample have data already published in the literature both in the X-ray band (Sazonov et al. 2005, Shinozaki et al. 2006) and in soft gamma-ray range (Beckmann et al. 2006, Molina et al. 2006). However, most of these works did not provide a detailed analysis of the broad band spectra of the sources in the sample nor attempted to put constraints on the high energy cut-off and reflection bump. The only exception is \\eso\\ analysed by our team using \\sax\\ and \\integral\\ data; here we use a recent \\xmm\\ observation of the source in conjunction with an updated \\integral\\ spectrum in order to have in the sample a well studied object to use as a source of reference. In Sect. \\ref{data} we present the data of our small sample while the models used to fit the spectra are described in Sect. \\ref{models}. Results and discussion are presented in Sect. \\ref{RD} and draw our conclusions. The single cases are discussed in more details in the appendix. ", "conclusions": "\\label{conclusions} We have presented spectral broad-band analysis of a small sample of seven absorbed Seyfert 2 selected at E$>$ 10 keV by \\integral/IBIS. The combined 0.2--200 keV spectra allowed us to perform a detailed study of the emission/absorption properties and to test possible correlations between different parameters (e.g. R vs $\\Gamma$ or $E_{c}$ vs $\\Gamma$); in particular, for the first time in these sources a measurement of the reflection components (both Compton bump and iron line) and high energy cut-off has been performed. An \\textit{a posteriori} check of our analysis procedure, which is affected by the limitation of using non simultaneous X-ray soft gamma-ray data, is performed providing for one well known source, \\eso, a direct comparison between our study and previous ones with \\sax\\ data. The values of $\\Gamma$, R and N$_{H}$ we found for this source are completely consistent with those obtained by Akylas et al. (2001) analysing simultaneous broad--band \\sax\\ observations. This evidence strengthens the validity of our fitting procedure and makes reliable our results. The main conclusions or our analysis are the followings. \\begin{itemize} \\item The average value of the absorbing column density is $\\bar{N}_H$ = (10.6$^{+0.2}_{-0.2}$) $\\times$ 10$^{22}$ \\cmM2, and the range observed suggests a Compton thin nature for all objects in the sample, that is a new results in the case of \\ngc, \\ic\\ and \\j104. This evidence is also confirmed by the ratio F$_x$/F$\\left [ \\rm OIII\\right ]$ $\\lambda$5007. \\item The continuum was reproduced with an e-folded power-law having a photon index in the range 1.3-1.9 and a high energy cut-offs well below 300 keV; this finding suggests that a cut-off is a common factor in Seyfert 2 galaxies with values in good agreement with those found in type 1 AGN. \\item We measure the Compton reflection component in five sources (\\120, \\ngc, \\eso, \\ic\\ and \\j104) and found an upper limit in two others (\\leda\\ and \\075). No evidence for a correlation between R and $\\Gamma$ is found. We also observed the iron line (in the case of \\chandra\\ spectra we found upper limits only on the equivalent width). Both reflection components are not immediately consistent with production in the cold absorbing gas identified in the molecular torus. Both the reflection fraction R and the equivalent width of the line are too high to be produced in the gas with the observed column density. A possible solution is that the absorption is more effective than the reflection, e.g. making the hypothesis that the absorbing/reflecting medium is not uniform, like a clumpy torus, or that the source is observed through a torus with a very shallow opening angle. \\item A soft excess component emerging below 2--3 keV is found in all sources with the only exception of \\075. This component is well reproduced by a thermal black body model having temperature in the range 0.2-0.9 keV, as typically observed in Sy 1. Alternatively a power-law with photon index equal to that of the illuminating continuum is also able to reproduce this excess is some cases; the ratio of soft versus primary continuum is of the order of a few per cent, as typically observed in Sy 2. A different case is that of \\120 which has, in the thermal model, a temperature kT=1.7$^{+0.1}_{-0.2}$ keV and, in the scattered scenario, A$_{IC}$/A$_{soft}$=0.12; both value are higher than observed in the other sources and, typically, in Seyfert. This strongly suggests that the source is characterized by a larger soft excess (possibly due to a starburst contribution) that is not present in the other sources. \\item \\075 is the only source in the sample that does not show any evidence for continuum excess at low energies. A narrow emission line around 2.5 keV is instead found in the \\chandra\\ data. \\end{itemize} We acknowledge the Italian Space Agency financial and programmatic support via contracts I/R/046/04 and I/023/05/0. We thank the referee for constructive suggestions." }, "0801/0801.2786_arXiv.txt": { "abstract": "The observed 511 keV line from the Galactic Bulge is a real challenge for theoretical astrophysics: despite a lot of suggested mechanisms, there is still no convincing explanation and the origin of the annihilated positrons remains unknown. Here we discuss the possibility that a population of slowly evaporating primordial black holes with the mass around $10^{16}-10^{17}$~g ejects (among other particles) low--energy positrons into the Galaxy. In addition to positrons, we have also calculated the spectrum and number density of photons and neutrinos produced by such black holes and found that the photons are potentially observable in the near future, while the neutrino flux is too weak and below the terrestrial and extra--terrestrial backgrounds. Depending on their mass distribution, such black holes could make a small fraction or the whole cosmological dark matter. ", "introduction": " ", "conclusions": "" }, "0801/0801.0260_arXiv.txt": { "abstract": "Magnetograms from the Vector SpectroMagnetograph (VSM) of the Synoptic Optical Long-term Investigations of the Sun (SOLIS) project are utilized to study the latitude distribution of magnetic flux elements as a function of latitude in the polar solar caps. We find that the density distribution of the magnetic flux normalized by the surface of the polar cap and averaged over months decreases close to the solar poles. This trend is more pronounced when considering only flux elements with relatively large size. The flux density of the latter is relatively flat from the edge of the polar cap up to latitudes of 70$^\\circ$--75$^\\circ$ and decreases significantly to the solar pole. The density of smaller flux features is more uniformly distributed although the decrease is still present but less pronounced. This result is important in studying meridional flows that bring the magnetic flux from lower to higher solar latitudes resulting in the solar cycle reversal. The results are also of importance in studying polar structures contributing to the fast solar wind, such as polar plumes. ", "introduction": "Polar regions of the Sun harbor numerous challenging solar phenomena, such as the source and acceleration process of the fast solar wind. Although it is widely believed that the magnetic field is main driver of most physical processes within the solar polar areas, these areas are not adequately characterized for observational and instrumental limitation reasons. Several solar phenomena (i.e., solar differential rotation, supergranular diffusion and meridional flows) couple together to transport poleward the mid-latitude magnetic flux of decaying active regions. This process leads to the formation of polar caps and their evolution through the solar cycle (see Babcock \\& Babcock 1955; Leighton 1964; DeVore \\& Sheeley 1987; Sheeley, Nash, \\& Wang 1987; etc.). Solar meridional flows have been shown, both observationally and theoretically, to be the main mechanism of zonal flux transport (Howard 1974; Durrant, Turner, \\& Wilson 2004; Duvall 1979; LaBonte \\& Howard 1982). However, the weak flow, of few times 10~m~s$^{-1}$ at best, is not easily measurable and direct measurements can not be achieved beyond solar mid-latitudes ($\\sim45^\\circ$). Thus, it is important to obtain additional constraints on these flows close to the solar poles, which would be of great use for solar dynamo and flux transport models. Raouafi, Harvey \\& Solanki (2006a,b; 2007) studied polar plumes EUV spectral emissions using different models. By comparing theoretical results to observations, they found that polar plumes would preferentially be based more than $10^\\circ$ away from the solar pole. Saito (1958) noticed that white-light plumes in eclipse observations were also rooted close to polar hole edges. The study of the polar flux latitude-distribution would confirm Raouafi et al.'s findings and constrain the magnetic flux transport (e.g., meridional flows) that drive such a distribution. ", "conclusions": "The high quality of magnetograms from SOLIS allowed us to characterize the flux concentration distribution as a function of latitude over a relatively large period of time. This was not possible earlier mainly because of instrumental limitations. The obtained results are important for studies of magnetic flux transport. They provide additional constraints on solar phenomena such as meridional circulation that is not possible to measure beyond $45^\\circ-50^\\circ$ nor how it functions near solar poles. Meridional flows, for instance, are an important input for solar dynamo and flux transport models. Our results on the density distribution of the magnetic flux concentrations at the polar regions suggests that the mechanisms responsible for the flux transport increasingly lose strength within the last 20 degree latitude before reaching the solar poles." }, "0801/0801.2579_arXiv.txt": { "abstract": "For extrasolar planets with orbital periods, P$>$10 days, radial velocity surveys find non-circular orbital eccentricities are common, $\\langle e\\rangle\\sim 0.3$. Future surveys for extrasolar planets using the transit technique will also have sensitivity to detect these longer period planets. Orbital eccentricity affects the detection of extrasolar planets using the transit technique in two opposing ways: an enhancement in the probability for the planet to transit near pericenter and a reduction in the detectability of the transit due to a shorter transit duration. For an eccentricity distribution matching the currently known extrasolar planets with P$>$10 day, the probability for the planet to transit is $\\sim 1.25$ times higher than the equivalent circular orbit and the average transit duration is $\\sim 0.88$ times shorter than the equivalent circular orbit. These two opposing effects nearly cancel for an idealized field transit survey with independent photometric measurements that are dominated by Poisson noise. The net effect is a modest $\\sim 4\\%$ increase in the transiting planet yield compared to assuming all planets have circular orbits. When intrinsic variability of the star or correlated photometric measurements are the dominant source of noise, the transit detectability is independent of the transit duration. In this case the transit yield is $\\sim$25\\% higher than that predicted under the assumption of circular orbits. Since the Kepler search for Earth-sized planets in the habitable zone of a Solar-type star is limited by intrinsic variability, the Kepler mission is expected to have a $\\sim$25\\% higher planet yield than that predicted for circular orbits if the Earth-sized planets have an orbital eccentricity distribution similar to the currently known Jupiter-mass planets. ", "introduction": "The known extrasolar planets possess a broad distribution of orbital eccentricity \\citep{BUT06} (see Figure~\\ref{fig:ecchist}). The short period, Hot Jupiter (P$<$10 day) planets predominately have circular orbits. However, at longer orbital periods circular orbits become a minority and the median eccentricity for extrasolar planets e$\\sim$0.3. At the extreme eccentricity end, there are three planets, HD 80606b \\citep{NAE01}, HD 20782b \\citep{JON06}, and HD 4113b \\citep{TAM07}, having e$>$0.9. HD 80606b comes closer to its stellar host (a=0.033 AU) than many of the circular orbit Hot Jupiter planets. \\citet{FOR07} recently reviewed the various mechanisms invoked to explain the distribution of eccentricities for the known extrasolar planets. Interactions with a stellar companion, planetary companion, passing star, gaseous disk, planetesimal disk, and stellar jets have all been proposed to modify the orbital eccentricity of extrasolar planets. Limited discussions in the literature have been given to the impact orbital eccentricity has on a transit survey for extrasolar planets. \\citet{TIN05} discuss the impact of eccentricity on their $\\eta$ parameter ($\\eta$ is the ratio of the observed transit duration to an estimate of the transit duration). As expected, they find transits occur near pericenter (apocenter) are shorter (longer) in duration than the circular orbit case, and they show that the transit duration of an eccentric is typically shorter than a circular orbit of the same period. However, their discussion was focused on the impact of orbital eccentricity on their $\\eta$ parameter. \\citet{MOU06} also discuss how the transit duration is affected by orbital eccentricity, but they do not quantify the impact this will have on transit surveys. Recently, \\citet{BAR07} derives the probability for a planet on an eccentric orbit to transit and conclude that the photometric precision of current surveys and future surveys, such as Kepler, is insufficient to determine the orbital eccentricity solely from the light curve. \\citet{BAR07} concludes that without knowledge of the eccentricity from radial velocity data or independent measurement of the stellar host radius, the habitability of planets detected with Kepler will remain unknown. Neglecting the impact eccentricity has on transit detections is justified for the current sample of transiting planets given the predominance of circular orbits for the Hot Jupiters and the strong bias of transit surveys against finding long period planets on circular orbits \\citep{GAU05}. The announced transit for the planet orbiting HD 17156\\footnote{First detected by the radial velocity technique \\citep{FIS07}} with a 21 day period and eccentric orbit \\citep{BARB07} is a precursor for the kinds of planets detectable in transit surveys. As transit surveys continue, longer period transiting planets may be discovered. More importantly, the recently launched COROT mission will surpass current surveys for sensitivity to longer period planets \\citep[P$\\sim$ several months; ][]{BORD03}. Also, the Kepler mission, scheduled for launch in 2009, has a goal to find Earth-sized objects at 1 AU from their host star \\citep{BORU04}. The main purpose of this paper is to show that given the distribution of eccentricities for the currently known extrasolar planets, eccentricity should not be ignored in assessing the detectability of transiting giant planets when the transit survey is sensitive to planets with P$>$ 10 day. In addition to longer period planets, the COROT and Kepler missions also will detect transiting planets with small, Earth-sized radii. The eccentricity distribution for planets less massive than the currently known Jupiter-mass planets is beginning to be explored. \\citet{RIB07} and \\citet{FOR07} provide tentative evidence for the tendency of lower mass planets to have lower eccentricities in the current sample of radial velocity planets. There is theoretical agreement that a proto-planetary gas disk strongly damps the eccentricity of non-gap opening embedded low-mass planets \\citep{GOL80, CRE07}. However, the theoretical models show that once the gas disk dissipates, dynamical interactions amongst the planets results in a random-walk diffusion that leads to an increasing eccentricity that takes on a Rayleigh-distribution similar to what is observed \\citep[dotted line in Figure~\\ref{fig:ecchist}; ][]{JUR07,ZHO07}. Opposing the increases in eccentricity from planet-planet scattering, late stage interactions with planetesimals can preferentially damp the eccentricities of lower mass planets enabling theoretical models to achieve the low eccentricities of the terrestrial planets of the Solar System \\citep{RAY06} and possibly explain the current trend of lower eccentricities for lower mass planets \\citep{FOR07}. Given the number of potential physical processes that can affect orbital eccentricity, it is premature to assume that the typical Earth-like planet has an eccentricity near zero like the Solar System. In this study, \\S~\\ref{sec:eccdist} reviews the observed eccentricity distribution of the known radial velocity extrasolar planets. The broad distribution of orbital eccentricity has two main effects on the sensitivity of a transit survey. First, the planet-host separation varies along an eccentric orbit, enhancing the probability to transit when the planet is relatively closer to the stellar host. \\S~\\ref{sec:tranprob} quantifies the net affect on the transit probability for a population of planets with non-circular orbits. Second, the planet velocity varies along and eccentric orbit, resulting in a reduction or lengthening of the transit duration. \\S~\\ref{sec:trandur} quantifies the distribution of transit durations resulting from a population of planets on eccentric orbits. \\S~\\ref{sec:disc} describes how the transit duration affects transit detection for transit surveys in the limit of various noise sources. \\S~\\ref{sec:conc} concludes by quantifying the net result of the two aforementioned effects, the enhanced probability to transit and the reduced detectability, on the yield from transit surveys. ", "conclusions": "\\label{sec:conc} Orbital eccentricity results in an enhanced probability for a planet to transit and potentially a reduction in the transit detectability. The overall yield from a transit survey is given by $N_{\\rm det}\\propto{\\rm Prob}_{{\\rm T}}\\times N_{\\rm obj}$. The results from this study can be used to scale the overall yield from a transit survey based on assuming all planets are on circular orbits for an assumed distribution of orbital eccentricity. The enhanced probability for a planet to transit, ${\\rm Prob}_{{\\rm T}}$ with a distribution of orbital eccentricity scaled to the circular orbit case is given by Equation~\\ref{eq:probtrane}. The reduced number of transiting planets detectable $N_{\\rm obj}$ scaled to the circular orbit case for an ideal transit survey where the photometric noise is white and dominated by Poisson error is given in \\S~\\ref{sec:disc}. Multiplying these two factors provides the overall yield of an ideal transit survey scaled to assuming all planets are on circular orbits. Figure~\\ref{fig:tranyield} show that these two opposing effects nearly cancel over the parameters that characterize the eccentricity distribution. Thus, for an idealized transit survey with the currently observed orbital eccentricity distribution, the overall yield will be 4\\% greater than assuming all planets are on circular orbits. The result for an idealized transit survey with a Rayleigh distribution of orbital eccentricities gives a similar enhancement (4\\%). \\begin{figure} \\includegraphics[scale=0.9,viewport=0 100 350 350]{f6.ps} \\caption{The overall yield from an idealized transit survey as function of the observe eccentricity distribution model parameters scaled to the assumption that all planets have circular orbits. The transit survey is assumed to have independent photometric measurements that are dominated by Poisson noise. The abscissa indicates $e_{\\rm max}$, and the curves are for selected values of $e_{\\rm crit}$ as labeled. For the model parameters shown in Figure~\\ref{fig:ecchist}, $N_{\\rm det}\\sim 4\\%$ higher than assuming all planets have circular orbits ({\\it dotted line}).\\label{fig:tranyield}} \\end{figure} In ground-based transit surveys, correlated measurements limit transit detectability \\citep{PON06} and intrinsic variability of the star will limit the detectability for Earth-sized planets for the Kepler mission \\citep{JEN02}. In both cases the transit detectability is independent of the transit duration, in which case $N_{\\rm obj}$ is independent of the orbital eccentricity. For these cases, the transit survey will have higher returns by a factor of $\\langle {\\rm Prob}_{{\\rm T}e}\\rangle$ (Equation~\\ref{eq:probtrane}) than estimated by assuming all planets are on circular orbits. However, the reduced planet yield in a transit survey due to intrinsic stellar variability or correlated measurements must be properly accounted for. If every dwarf star has an Earth-sized planet orbiting in the habitable zone, then assuming circular orbits, the Kepler mission expects to detect 100 Earth-sized planets in the habitable zone \\citep{BORU04}. The work presented here indicates Kepler will have $\\langle {\\rm Prob}_{{\\rm T}e}\\rangle\\sim 25\\%$ higher yield if Earth-sized planets in the habitable zone have a planet eccentricity distribution similar to the currently known sample of giant planets from radial velocity surveys. However, if Earth-sized planets with e$\\sim$0.9 are as common as e$\\sim$0 then, the yield of Earth-sized planets could be 80\\% higher from the Kepler mission assuming the high-e, short transit duration planets are still detectable. The dependence of the transit survey yield on the uncertain underlying orbital eccentricity distribution implies an uncertainty in measuring the frequency of terrestrial planets in the habitable zone \\citep[a major goal of the Kepler mission][]{BORU04} . An analysis of the transit yield from a transit survey that assumes all planets are on circular orbits will overestimate the frequency of habitable planets if high eccentricities are common and not taken into account. In practice a variety of noise regimes affect a transit survey and accurate yields necessitate an accurate understanding of the photometric noise, stellar sample, and underlying eccentricity distribution \\citep{BUR06,GOU06,FRE07}." }, "0801/0801.0110_arXiv.txt": { "abstract": "We explore the abundance of light clusters in core-collapse supernovae at post-bounce stage in a quantum statistical approach. Adopting the profile of a supernova core from detailed numerical simulations, we study the distribution of light bound clusters up to alpha particles (2 $\\leq A \\leq$ 4) as well as heavy nuclei ($A > 4$) in dense matter at finite temperature. Within the frame of a cluster-mean field approach, the abundances of light clusters are evaluated accounting for self-energy, Pauli blocking and effects of continuum correlations. We find that deuterons and tritons, in addition to $^3$He and $^4$He, appear abundantly in a wide region from the surface of the proto-neutron star to the position of the shock wave. The appearance of light clusters may modify the neutrino emission in the cooling region and the neutrino absorption in the heating region, and, thereby, influence the supernova mechanism. ", "introduction": " ", "conclusions": "" }, "0801/0801.3219_arXiv.txt": { "abstract": "{Amino acids are building blocks of proteins and therefore key ingredients for the origin of life. The simplest amino acid, glycine (NH$_2$CH$_2$COOH), has long been searched for in the interstellar medium but has not been unambiguously detected so far. At the same time, more and more complex molecules have been newly found toward the prolific Galactic center source Sagittarius B2.} {Since the search for glycine has turned out to be extremely difficult, we aimed at detecting a chemically related species (possibly a direct precursor), amino acetonitrile (NH$_2$CH$_2$CN).} {With the IRAM 30m telescope we carried out a complete line survey of the hot core regions Sgr~B2(N) and (M) in the 3\\,mm range, plus partial surveys at 2 and 1.3\\,mm. We analyzed our 30m line survey in the LTE approximation and modeled the emission of all known molecules simultaneously. We identified spectral features at the frequencies predicted for amino acetonitrile lines having intensities compatible with a unique rotation temperature. We also used the Very Large Array to look for cold, extended emission from amino acetonitrile.} {We detected amino acetonitrile in Sgr~B2(N) in our 30m telescope line survey and conducted confirmatory observations of selected lines with the IRAM Plateau de Bure and the Australia Telescope Compact Array interferometers. The emission arises from a known hot core, the Large Molecule Heimat, and is compact with a source diameter of 2$\\arcsec$ (0.08 pc). We derived a column density of $2.8 \\times 10^{16}$ cm$^{-2}$, a temperature of 100 K, and a linewidth of 7 km~s$^{-1}$. Based on the simultaneously observed continuum emission, we calculated a density of $1.7 \\times 10^8$ cm$^{-3}$, a mass of 2340 M$_\\odot$, and an amino acetonitrile fractional abundance of $2.2 \\times 10^{-9}$. The high abundance and temperature may indicate that amino acetonitrile is formed by grain surface chemistry. We did not detect any hot, compact amino acetonitrile emission toward Sgr~B2(M) or any cold, extended emission toward Sgr~B2, with column-density upper limits of $6 \\times 10^{15}$ and $3 \\times 10^{12-14}$ cm$^{-2}$, respectively.} {Based on our amino acetonitrile detection toward Sgr~B2(N) and a comparison to the pair methylcyanide/acetic acid both detected in this source, we suggest that the column density of both glycine conformers in Sgr~B2(N) is well below the best upper limits published recently by other authors, and probably below the confusion limit in the 1-3\\,mm range.} ", "introduction": "\\label{s:intro} Among the still growing list of complex molecules found in the interstellar medium, so-called ``bio''molecules garner special attention. In particular, the quest for interstellar amino acids, building blocks of proteins, has engaged radio and millimeter wavelength astronomers for a long time. Numerous published and unpublished searches have been made for interstellar glycine, the simplest amino acid \\citep{Brown79,Hollis80,Berulis85,Combes96,Ceccarelli00,Hollis03,Jones07,Cunningham07}. Its recent ``detection'' claimed by \\citet{Kuan03} has been persuasively rebutted by \\citet{Snyder05}. Since the early days of molecular radio astronomy, Sagittarius B2 has been a favorite target in searches for complex molecules in space. \\subsection{The target: Sagittarius B2} \\label{ss:sgrb2intro} \\object{Sagittarius B2} (hereafter Sgr~B2 for short) is a very massive (several million solar masses) and extremely active region of high-mass star formation at a projected distance of $\\sim 100$ pc from the Galactic center. Its distance from the Sun is assumed to be the same as the Galactic center distance, $R_0$. \\citet{Reid93}, reviewing various methods to determine $R_0$, arrived at a ``best estimate'' of 8.0 $\\pm$ 0.5 kpc, a value that we adopt in this article. It is supported by recent modeling of trajectories of stars orbiting the central black hole, which yields 7.94 $\\pm$ 0.42 kpc \\citep{Eisenhauer03}. There are two major centers of activity, \\object{Sgr~B2(M)} and \\object{Sgr~B2(N)} separated by $\\sim 2$ pc. In each of them, recent star formation manifests itself in a multitude of H{\\sc ii} regions of many sizes, from hypercompact to compact \\citep[][]{Gaume95}, and there is abundant material to form new stars evident by massive sources of molecular line and submillimeter continuum emission from dust \\citep{Lis91, Lis93}. \\subsubsection{Sgr B2 as part of the Central Molecular Zone} \\label{sss:sgrb2cmz} Some of the first detections of interstellar organic molecules (at cm-wavelengths!) were made toward Sgr~B2 \\citep[see][ for a historical perspective]{Menten04}. The low intrinsic line strengths make these cm lines unlikely candidates for detection. However, the situation is helped, first, by the fact that many of the transitions in question may have inverted levels \\citep{Menten04} and amplify background radio continuum emission which is very intense at cm wavelengths \\citep{Hollis07}. Second, the spatial distributions of many species are characterized by spatially extended emission covering areas beyond Sgr~B2 itself, filling single dish telescope beams, thus producing appreciable intensity even when observed with low spatial resolution \\citep[][]{Cummins86,Jones08}. This emission is characterized by low rotation temperatures, favoring lower frequency lines. Recent identifications of ``new'' species include glycolaldehyde CH$_2$OHCHO \\citep[][]{Hollis00,Hollis01}, ethylene glycol HOCH$_2$CH$_2$OH \\citep[][]{Hollis02}, and vinyl alcohol CH$_2$CHOH \\citep[][]{Turner01}. Sgr B2 and its surroundings are part of the Central Molecular Zone (CMZ) of our Galaxy, a $\\sim \\pm 0\\fdg3$ latitude wide band stretching around the Galactic center from longitude $l \\sim +1\\fdg6$ to $-1\\fdg1$ \\citep[see, e.g.,][]{Morris96}. The CMZ contains spatially extended emission of many complex organic molecules \\citep{Minh92,Dahmen97,Menten04,RequenaTorres06}. \\subsubsection{The Large Molecule Heimat} \\label{sss:introlmh} Near Sgr~B2(N), there is a hot, dense compact source that has a mm-wavelength line density second to no other known object. This source, for which \\citet{Snyder94} coined the name ``Large Molecule Heimat'' (LMH), is characterized by very high densities ($> 10^7$~cm$^{-3}$) and gas temperatures ($> 100$~K). In recent years arcsecond resolution interferometry with the BIMA array has resulted in the detection and imaging of increasingly complex organic species toward the LMH, such as vinyl cyanide CH$_2$CHCN, methyl formate HCOOCH$_3$, and ethyl cyanide CH$_3$CH$_2$CN \\citep[][]{Miao95,Miao97}, formamide NH$_2$CHO, isocyanic acid HNCO, and methyl formate HCOOCH$_3$ \\citep[][]{Kuan96a}, acetic acid CH$_3$COOH \\citep[][]{Mehringer97b,Remijan02}, formic acid HCOOH \\citep[][]{Liu01}, and acetone (CH$_3$)$_2$CO \\citep[][]{Snyder02}. All the interferometric observations are consistent with a compact ($< $few arcsec diameter) source that had already been identified as the source of high-density-tracing non-metastable ammonia line emission by \\citet{Vogel87} and thermal methanol emission by \\citet[][, their source ``i'']{Mehringer97a}. The LMH also hosts a powerful H$_2$O maser region \\citep[][]{Reid88}, which provides evidence that it is very young (see Sect.~\\ref{ss:sgrb2n}). \\subsection{The complex spectra of complex molecules} Complex molecules in general have large partition functions, in particular for the elevated temperatures ($> 100$ K) in molecular hot cores, dense and compact cloud condensations internally heated by a deeply embedded, young high-mass (proto)stellar object. Therefore, most individual spectral lines are weak and might easily get hidden in the ``line forest'' found toward these frequently extremely line-rich sources. To a large part, this forest consists of rotational lines, many of them presently unidentifiable, from within relatively low-lying vibrational states of molecules. Most of the candidate molecules from which these lines originate are known to exist in these sources, but laboratory spectroscopy is presently lacking for lines from the states in question. At this point in the game, unequivocally identifying a species in a spectrum of a hot core covering a wide spectral range requires the following steps: as described in detail in Sect.~\\ref{ss:modeling30m}, assuming Local Thermodynamic Equilibrium (LTE) (which applies at the high densities in hot cores) a model spectrum is calculated for an assumed rotation temperature, column density, line width and other parameters. This predicts lines of a given intensity at all the known frequencies. Then at least two conditions have to be fulfilled: (i) All predicted lines should have a counterpart in the observed spectrum with the right intensity and width -- no single line should be missing. (ii) Follow-up observations with interferometers have to prove whether all lines from the candidate species are emitted from the same spot. Given the chemical variety in hot core regions, this is a powerful constraint. Moreover, interferometer images tend to have less line confusion, since many lines that are blended in larger beam single-dish spectra arise from different locations or are emitted by an extended region that is spatially filtered out. Using an interferometer for aiding molecule identifications was pioneered by L. Snyder and collaborators who (mostly) used the Berkeley-Illinois-Maryland-Array (BIMA) to clearly identify a number of species in the Sgr~B2(N) Large Molecule Heimat (see Sect.~\\ref{sss:introlmh}). We carried out a complete line survey of the hot core regions Sgr~B2(N) and (M) with the IRAM 30m telescope at 3\\,mm, along with partial surveys at 2 and 1.3\\,mm. One of the overall goals of our survey was to better characterize the molecular content of both regions. It also allows searches for ``new'' species once we have identified the lines emitted by known molecules (including vibrationally and torsionally excited states). In particular, many complex molecules have enough lines in the covered frequency ranges to apply criterion (i) above. Once a species fulfils this criterion, interferometer measurements of selected lines can be made to check criterion (ii). \\subsection{Amino acetonitrile} One of our target molecules was amino acetonitrile (NH$_2$CH$_2$CN), a molecule chemically related to glycine. Whether it is a precursor to the latter is under debate (see Sect.~\\ref{ss:precursor}). Not many astronomical searches for amino acetonitrile have been reported in the literature. In his dissertation, \\citet[][]{Storey76} reported searches for the $J_{K_a,K_c} = 2_{11}-2_{12}$ and $1_{01}-0_{00}$ transitions at 1350.5 and 9071.7 MHz, respectively with the Parkes 64\\,m telescope. On afterthought, the only chance of success for their observations would have been if amino acetonitrile existed on large spatial scales, similar to the molecules described in Sect.~\\ref{sss:sgrb2cmz} (see Sect.~\\ref{ss:aanvla} for further limits on extended amino acetonitrile emission). Recently, \\citet{Wirstroem07} reported unsuccessful searches of a number of mm-wavelength transitions of amino acetonitrile toward a number of hot cores. Here, we report our detection of warm compact emission from amino acetonitrile in Sgr~B2(N) with the IRAM 30m telescope, the Plateau de Bure Interferometer (PdBI) and the Australia Telescope Compact Array (ATCA), and upper limits on cold, spatially extended emission from amino acetonitrile that we obtained with the NRAO Very Large Array (VLA). Section~\\ref{s:obs} summarizes the observational details. We present our results in Sect.~\\ref{s:results}. Implications in terms of interstellar chemistry are discussed in Sect.~\\ref{s:discussion}. Our conclusions are summarized in Sect.~\\ref{s:concl}. ", "conclusions": "" }, "0801/0801.3733_arXiv.txt": { "abstract": "We aim to understand cloud formation in substellar objects. We combined the non-equilibrium, stationary cloud model of Helling, Woitke \\& Thi (2008; seed formation, growth, evaporation, gravitational settling, element conservation) with the general-purpose model atmosphere code {{\\sc Phoenix}} (radiative transfer, hydrostatic equilibrium, mixing length theory, chemical equilibrium) in order to consistently calculate cloud formation and radiative transfer with their feedback on convection and gas phase depletion. We calculate the complete 1D model atmosphere structure and the chemical details of the cloud layers. The {{\\sc Drift-Phoenix}} models enable the first stellar atmosphere simulation that is based on the actual cloud {{\\it formation}} process. The resulting $(T,p)$ profiles differ considerably from the previous limiting {{\\sc Phoenix}} cases {{\\sc Dusty}} and {{\\sc Cond}}. A tentative comparison with observations demonstrates that the determination of effective temperatures based on simple cloud models has to be applied with care. Based on our new models, we suggest a mean T$_{\\rm eff}=1800$K for the L\\,-\\,dwarf twin-binary system \\objectname{DENIS J0205-1159} which is up to 500K hotter than suggested in the literature. We show transition spectra for gas-giant planets which form dust clouds in their atmospheres and evaluate photometric fluxes for a \\objectname{WASP-1} type system. ", "introduction": "Today's most efficient tools to interpret the observed spectra of substellar objects, i.e., brown dwarfs and planets, are 1D static atmosphere simulations. Comparisons with observations, based on the solution of the radiative-transfer problem which is done in great detail with respect to the gas-phase opacities (Tsuji 2002, 2005; Allard et al. 2001, Ackerman \\& Marley 2001, Burrows \\& Sharp 1999, Barman et al. 2005), ideally yield insight into the atmospheric structure and chemistry providing finger prints of its evolutionary state. Brown dwarf and planetary atmospheres have a far more complex chemistry than stellar objects due to the formation of clouds, which bind chemical elements and, hence, strongly influences the remnant gas phase inside the atmosphere. The presence of such clouds was previously simplified in static model atmosphere codes. We now move a significant step forward by kinetically treating the chemistry of cloud formation as a phase-non-equilibrium process in the framework of model atmosphere simulations. Our dust model has been studied so far for a given $(T, p, v_{\\rm conv})$ structure ($T$ - gas temperature, $p$ - gas pressure, $v_{\\rm conv}$ - convective velocity), and we present in this letter the first consistent simulation of cloud micro-physics and atmospheric structure which allows for the first time to study the feedback of the dust {\\it formation} onto the atmospheric structure. We present our first consistent {\\sc Drift-Phoenix} results together with a tentative comparison to the observed spectrum of \\objectname{ DENIS J0205--1159} and point out possible uncertainties in present T$_{\\rm eff}$ determinations (Sect.~\\ref{s:appl}). We, however, leave aside the issue of hydrodynamical cloud formation (e.g. Showman et al. 2006, Knutson et al. 2007, Rauscher et al. 2007) which unavoidably has to deal with the turbulent closure problem (Helling et al. 2004, Helling 2007). We calculate transition spectra for gas-giant planets and a {\\sc WASP}-1 type star, and evaluate the photometric fluxes for the 2MASS system, for the VISIR system, and for the IRAC band systems. ", "conclusions": "The modelling of cloud formation in substellar atmospheres has a strong impact on the objects temperature-pressure structure which in turn determines the cloud's chemical composition, the grain size distribution function, and the cloud's location inside the atmosphere. These are results of our consistent simulation of substellar atmospheres and detailed non-equilibrium dust cloud formation. We demonstrate synthetic transition spectra for gas-giant planets and calculate synthetic photometric fluxes based on our consistent solution of dust cloud formation and radiative transfer problem. A future goal is to extend our model to solar-system-like planets with much cooler atmosphere. As we have shown for the field object {\\sc Denis} J0205-1159, stellar parameter determinations based on a comparison with synthetic spectra can vary considerable. As one consequence of this, we have set out to conduct a component-based study \\footnote{http://phoenix.hs.uni-hamburg.de/BrownDwarfsToPlanets1/ } where e.g. cloud compositions, dust-to-gas-ratios, and grain sizes are compared for different cloud models (Helling et al. 2007, 2008)." }, "0801/0801.4325_arXiv.txt": { "abstract": "We analyze the hitherto available space-based X-ray data as well as ground-based optical data of the X-ray transient 080109/SN\\,2008D. From the data we suggest that ({\\it i}) The initial transient ($\\lesssim 800$ sec) is attributed to the reverse shock emission of a mildly relativistic ($\\Gamma \\sim$ a few) outflow stalled by the dense stellar wind. ({\\it ii}) The subsequent X-ray afterglow ($\\lesssim 2\\times 10^4$ sec) can be ascribed to the forward shock emission of the outflow, with a kinetic energy $\\sim 10^{46}$ erg, when sweeping up the stellar wind medium. ({\\it iii}) The late X-ray flattening ($\\gtrsim 2\\times 10^4$ sec) is powered by the fastest non-decelerated component of SN\\,2008D's ejecta. % ({\\it iv}) The local event rate of X-ray transient has a lower limit of $\\sim 1.6\\times 10^4~{\\rm yr^{-1}~Gpc^{-3}}$, indicating a vast majority of X-ray transients have a wide opening angle of $\\gtrsim 100^\\circ$. ({\\it v}) Transient 080109/SN\\,2008D indicates a continuum from GRB-SN to under-luminous GRB-/XRF-SN to X-ray transient-SN and to ordinary Ibc SN (if not every Ibc SN has a relativistic jet), as shown in Figure 2 of this {\\it Letter}. ", "introduction": "\\label{sec:Into} During the past decade, long-duration ($\\gtrsim 2$sec) $\\gamma-$ray bursts (GRBs), including the subclass of X-ray flashes (XRFs), have been found (1) to be driven by the core-collapse of massive stars \\cite{Woosley93}; thus (2) to be associated with a rare variety ($\\sim1\\%$) of type Ibc supernovae (SNe), the so-called hypernovae (HN) \\cite{Galama98,Hjorth03,Stanek03,Male04,Sollerman06,Camp06} (but also see Fynbo et al. 2006); and (3) in general to be hosted by the star-forming dwarf galaxies with low metallicity \\cite{Fyn03,Fru06,Stanek06}. Though the association of GRB/XRF and Ibc SN has been pinned down, what channels make a dying star to produce a GRB or an XRF, and not just a Ibc SN, is still unclear. The progenitor's mass, metallicity, angular momentum, and the configuration and strength of its internal magnetic field play important roles for the generation of GRBs/XRFs and ordinary Ibc SNe. The serendipitous discovery of the X-ray transient 080109/SN\\,2008D may shed light on filling in this gap between energetic GRBs/XRFs and ordinary Ibc SNe. We will analyze space- and ground-based data of this transient and SN, focusing on X-ray/radio data because observationally they trace the fastest component of the transient/SN outflow while optical data trace the slower SN ejecta (e.g., Soderberg et al. 2006). ", "conclusions": "Transient 080109/SN\\,2008D presents the first evidence for a mild-relativistic outburst, $\\sim 10^{46}\\,{\\rm erg}$, preceding the main SN component, thus confirming previous speculation in SN\\,2005bf (Folatelli et al. 2006). It sets a lower limit of the local event rate of its kind as $1/3{\\rm{yr}}/({\\rm{0}}{\\rm{.0276 Gpc}})^3 \\sim 1.6 \\times 10^4 {\\rm{yr}}^{{\\rm{ - 1}}} {\\rm{Gpc}}^{{\\rm{ - 3}}}$, comparable with the local rate of Ibc SNe, $\\sim 4.8 \\times 10^4 {\\rm{yr}}^{{\\rm{ - 1}}} {\\rm{Gpc}}^{{\\rm{ - 3}}}$, and thus indicates a vast majority of X-ray transients have a wide opening angle of $\\gtrsim 100^\\circ$. The collimation-corrected energy is of $\\sim 5\\times 10^{45}$ erg. The wide angle budget, together with the self-consistent interpretation of the transient and its early afterglow with the on-beam model, largely rules out the off-axis viewing model for this transient. The host NGC2770, a spiral galaxy with copious ${\\rm H\\alpha}$ sign, stands out from the star-forming dwarf galaxies typically hosting GRBs/XRFs. Transient 080109 puts itself on the upper border of the nearby GRB/XRF collection in terms of the host metallicity (Berger \\& Soderberg 2008b; Sollerman et al. 2005). As a result, this event may unveil a continuum from energetic GRB (top-right of Figure 2) to ordinary Ibc SN (bottom-left of Figure 2). (1) Whether or not every Ibc SN has a quasi-jet outburst proceeding the main SN component is still uncertain even the discovery of Transient 080109. For this reason we mark the ordinary Ibc SN and X-ray transient/SN populations with a dash ellipse in Figure 2. (2) Soderberg et al. (2006) showed that producing GRBs/XRFs needs a relativistic ejecta carrying at least $10^{48}$ erg. We show in this {\\it Letter} that X-ray transient population couples $\\sim 10^{46}$ erg to relativistic material regarding this found one marks the transition between GRB/XRF and ordinary Ibc. (3) While under-luminous GRBs/XRFs are likely powered by moderate-relativistic material, X-ray transients are likely powered by mild-relativistic material. (4) GRBs have an average opening angle of $\\sim \\ 10^\\circ$ while a vast majority (if not all) of X-ray transients have a much wider one of $\\sim 100^\\circ$. There is a negative correlation between radiated energy and opening angle from GRB to XRF to X-ray transient. (5) Materials with higher bulk Lorentz factor tend to have a shallower energy-velocity distribution leading to spikeful behavior as shown in GRB/XRF prompt lightcurves and hypernova's broad-lined spectra. Materials with lower bulk Lorentz factor tend to have a steeper energy-velocity distribution and thus largely couple with each other leading to spikeless/little-spiked behavior as shown in various optical afterglows. The decay laws in terms of velocity for each event in Figure 2 (from left to right) matches this principle." }, "0801/0801.4864_arXiv.txt": { "abstract": "We studied the radio structure of high-redshift ($z>3$) quasars with VSOP at 1.6 and 5~GHz. These sources are the most distant objects ever observed with Space VLBI, at rest-frame frequencies up to $\\sim25$~GHz. Here we give an account of the observations and briefly highlight the most interesting cases and results. These observations allowed us, among other things, to estimate the mass of the central black holes powering these quasars, to identify large misalignments between the milli-arcsecond (mas) and sub-mas scale radio structures, and to detect apparent superluminal motion at sub-mas scale. ", "introduction": "Twenty of the most distant ($z>3$) radio quasars have been proposed for Space Very Long Baseline Interferometry (SVLBI) observations at either 1.6 or 5~GHz or both in the VLBI Space Observatory Programme \\citep[VSOP;][]{hira00}. These were the brightest known, with total flux densities higher than $\\sim400$~mJy. Our goal was to study the structural properties of high-redshift sources at the highest possible resolution at emitted frequencies up to $\\sim20-25$~GHz. This allowed us to investigate source compactness, relativistic jet propagation, and eventually to estimate physical parameters of the jets and the active nuclei feeding them. ", "conclusions": "" }, "0801/0801.3169_arXiv.txt": { "abstract": "We propose a model of allocating galaxies in cosmological N-body simulations. We identify each subhalo with a galaxy, and assign luminosity and morphological type assuming that the galaxy luminosity is a monotonic function of its host subhalo mass. The morphology assignment is made by using two simple relations between subhalo mass and galaxy luminosity of different types. One is using a constant ratio in luminosity of early (E/SO) and late (S/Irr) type galaxies at a fixed subhalo mass. And the other assumes that galaxies of different morphological types but having an equal luminosity have a constant ratio in their subhalo masses. We made a series of comparisons of the properties of these simulated galaxies with those of the SDSS galaxies. The resulting simulated galaxy sample is found to successfully reproduce the observed local number density distribution except for in high density regions. The luminosity function is studied as a function of local density. It was found that the observed luminosity functions in different local density environments are overall well-reproduced by the simulated galaxies. Discrepancy is found at the bright end of the luminosity function of early types in the underdense regions and at the faint end of both morphological types in very high density regions. A significant fraction of the observed early type galaxies in voids seems to have undergone a relatively recent star formation and became brighter. The lack of faint simulated galaxies in dense regions may be due to the strong tidal force of the central halo which destroys less massive satellite subhalos around in the simulation. The mass-to-light ratio is found to depend on the local density in the way similar to that observed in the SDSS sample. We have found an impressive agreement between out simulated galaxies and the SDSS galaxies in the dependence of the central velocity dispersion on the local density and luminosity. ", "introduction": "The current galaxy formation paradigm can be characterized by ``hierarchical clustering''. This means that massive dark matter halos form by merging less massive halos and/or by accreting ambient matter, and also that a dark matter halo governs the evolution of the galaxy residing inside. Most galaxies are believed to be hosted by the dark matter halos becuase the halos can provide a deep potential well for baryonic matter to condense and cool down. This leads to triggering of the star formation; gas sufficiently accumulated in the halo potential center begins to experience many hydrodynamic processes, such as radiative cooling, star-formation, supernova explosion, and chemical enrichments. All of these processes play an important role in making the visible galaxies in the end. Because galaxies as a building block of the large scale structures consequently follows the evolution of their host halos over the cosmic history, understanding gravitational evolution of dark halos is very important for the study of galaxy formation and evolution. Over the past few decades, cosmological simulations have been proved useful for the study of structure formation. Simulations have boosted up many investigations of the nonlinear structure evolutions and their results have been extensively compared to the observations in various aspects of interest. Many studies have reported successful recovery of observational features of the galaxy distribution like the two-point correlations of galaxies \\citep{conroy06,kravtsov04,berlind02}, topology \\citep{park05a,park05b,gott06}, and the environmental dependence of spin distributions \\citep{cervantes-sodi07}. Detailed modeling of galaxy evolution and understanding of the galaxy properties have been possible with the recent advent of the huge redshift surveys such as Sloan Digital Sky Survey\\footnote{http://www.sdss.org} (SDSS) and 2dFGRS\\footnote{http://www.aao.gov.au/2df/}. These larger surveys provide a rich information on the formation and evolution of galaxy. To meet a requirement to establish a relation between simulated structures and observed galaxies many techniques have been introduced. The semi-analytic model(SAM) of galaxy formation \\citep{cole94,kauffmann97,baugh06} is based on the hierarchical clustering of halos whose merging history trees are built by generating a set of Gaussian random numbers. The mass growth history of halos on various mass scales can be generated and traced by this method. Numerous SAM parameters are implemented in the merging history to reflect the hydrodynamical processes of heating, cooling, star formation, and aging the stellar populations. Their parameter values are fine tunned for the best description of collective properties of observed galaxies. But as the number of observational contraints increases, the number of parameters increases and the SAM becomes much complicated. Hydrosimulations employ direct formulations of the hydrodynamic processes (\\citealt{weinberg06}). Two types of hydro simulations are now widely adopted; the Lagrangian \\citep{monaghan92,hernquist89} and Eulerian methods (\\citealt{tasker06,harten97}, and for comparative study, see \\citealt{heitmann05,thacker00}). They set gas particles or gas grids in the system of interest. Then, gas dynamics are taken into account by calculating the hydrodynamic interactions between neighbor particles or by solving differential hydro equations between adjacent grids. This method has several weak points; it suffers from the lack of resolutions in space and mass as the N-body simulations. And in most cases, the star formation and supernova explosions are far beyond the simulation resolution. Some processes such as radiative transfer, star formation, supernovae feedback, and initial stellar mass function, are not thoroughly understood and are difficult to parameterize in a well established way. One of the offsprings of the SAM is the Halo Occupation Distribution (HOD; \\citealt{zheng05,seljak00,berlind02}) model. It relates the Friend-of-Friend (FoF) halo mass to the number of contained subhalos (or galaxies) using a conditional probability, $P(N|M)$, where $M$ is the FoF halo mass and $N$ is the number of subhalos inside the FoF halo. This probability function is obtained from numerical simulations \\citep{kravtsov04,berlind02,jing98} or from observations \\citep{zehavi04,abazajian05,zheng07} fitting the model to the observed two-point correlation functions. Another variant of the SAM is a one that adopts the one-to-one monotonic correspondence between the galaxy luminosity and subhalo mass \\citep{marinoni02,vale04,vale06,shankar06}. This method is often called the ``correspondence model''. It is simpler than other methods because it requires only two prerequisites; the luminosity function of galaxies and mass function of subhalos. Its key assumption that a more massive subhalo hosts a brighter galaxy, is consistent with the hierarchical clustering picture. As the halo mass grows, its luminosity is expected to grow in general because merging of halos may be followed by the merging of galaxies. By matching these two functions, the subhalo mass is mapped to galaxy luminosity. However, sometimes a galaxy may survive the merger event because baryonic component of a galaxy is usually more concentrated and is more tightly bound than its dark counterpart. These ``orphan'' galaxies \\citep{gao04} which have no separate corresponding halos could exist in the cluster regions. Also halos of mass below a certain characteristic scale are unable to provide baryonic matter with gravitational attraction strong enough to resist against supernova explosions which can tear up the small-mass system. Even though such cases are many, it is sound to expect that the galaxy census is closely related to the population of subhalos because most of the observed galaxies are field galaxies who have their own dark halos. It is worth investigating the hypothesis of the subhalo-to-galaxy correspondence model of galaxy formation. In this paper, we apply the subhalo-galaxy correspondence model to a large N-body simulation and compare the statistical properties of simulated galaxies with those of galaxies observed by the SDSS. We organize this paper as follows. In section 2 we describe our simulation and the subhalo finding method. In section 3, we implement the subhalo-galaxy correspondence model and show how to assign morphological types to mock galaxies. In section 4, the local density is introduced to quantify local environments and the local density distributions of mock and SDSS galaxies are investigated. We also compare the luminosity distribution of simulated galaxies with those of the SDSS galaxies in section 5. The environmental dependence of central velocity distributions are studied for the mock and SDSS galaxies in section 6. And discussions and conclusions follow in section 7. ", "conclusions": "\\label{summary} We have proposed a model to assign galaxy luminosity and morphology to the dark subhalos directly identified in cosmological N-body simulations. It is assumed that the galaxy luminosity is a monotonic function of its host halo mass. In the $\\kappa$ model we assume that the halo masses of early and late type galaxies of equal luminosity have a constant ratio. Another alternative model is the $\\beta$ model which assumes the constant luminosity ratio between equal-mass galaxies of different types. This model has been proposed by \\citet{marinoni02} who found that the observed $B$--band luminosity function of galaxies can be reproduced from the PS function assuming double power-law mass-to-light ratios and derived halo occupation numbers. It has been expanded by \\citet{vale04,vale06} who adopted the satellite halo mass functions and directly link subhalos to the observed galaxies. In this paper, we have introduced the ratio of luminosity of the early and late morphological type galaxies at a given halo mass, and derived type-specific mass-to-light ratios as a function of subhalo mass. They are used to assign luminosity and morphology to subhalos. The mass-to-light ratio of the early type galaxies derived in this way has a minimum value of $\\Upsilon_u\\simeq100$ at the scale of $M_u\\simeq3\\times10^{11}h^{-1}{\\rm M_\\odot}$ in the $\\beta=2$ model. The mass-to-light ratio starts to exponentionaly increase below $M_u$ and increases in a power law above $M_u$. We use the large-scale background galaxy number density as an environmental parameter. The smooth galaxy density field is obtained by using the adaptive spline kernel which enabled us to resolve crowded regions well. The local density distribution of the SDSS galaxies is well described by the log normal function. The obtained log normal parameter values indicate that brighter galaxies tend to locate in dense regions. The local density distribution of the simulated galaxies is quite similar to that of the SDSS galaixes in voids and moderate-density regions for both morphological types. The underestimation of galaxy population in clusters is thought to be due to the evaporation of subhalos by the dynamical friction and tidal stripping. This can also explain the discrepance in the luminosity function of the early type galaxies in high density regions. Recently, \\citet{gott06} measured the genus statistic from a large sample of the SDSS galaxies and compared it with those of mock galaxies created by the three distinctive methods such as the semi-analytic galaxy formation model applied to the Millennium run \\citep{springel05}, a hydrodynamic simulation, and the subhalo-galaxy correspondence model adopted in this paper. It was found that the observed topology of large scale structure was best reproduced by the subhalo-galaxy correspondence model even though other models were also consistent with the observation. However, the observed topology was marginally inconsistent with all simulations in the sense that it showed a strong meatball topology at the significance level of $2.5 \\sigma$ at the scale studied. The prominence of the isolated high density regions in the observation seems to be due to the Sloan Great Wall which was a dominant structure in the sample analyzed. We have found an impressive agreement between our simulated galaxies and the SDSS galaxies in the dependence of the central velocity dispersion on the local density and luminosity. The early type galaxies tend to have higher $\\sigma_v$ or higher mass in high density regions at a given luminosity when they are brighter than about $\\mathcal{M}_*$. In other words, these bright galaxies tend to become relatively fainter in high density regions at a given halo mass. This interesting dependence of the mass-to-light ratio on environment was successfully reproduced by our subhalo-galaxy correspondence model of galaxy formation. A more detailed study of this phenomenon will be presented in a forthcoming paper." }, "0801/0801.4055_arXiv.txt": { "abstract": "{ A brief review of various methods to calculate radiative accelerations for stellar evolution and an analysis of their limitations are followed by applications to Pop I and Pop II stars. Recent applications to Horizontal Branch (HB) star evolution are also described. It is shown that models including atomic diffusion satisfy Schwarzschild's criterion on the interior side of the core boundary on the HB without the introduction of overshooting. Using stellar evolution models starting on the Main Sequence and calculated throughout evolution with atomic diffusion, radiative accelerations are shown to lead to abundance anomalies similar to those observed on the HB of M15. ", "introduction": "In his book, \\citet{Eddington26} evaluated the equilibrium concentrations to be expected if atomic diffusion in the presence of differential radiation pressure were efficient. He concluded that some heavy metals should then completely dominate the spectrum. Since light and heavy elements are present in stellar spectra, he concluded (\\S 193, 196) that some mixing process made atomic diffusion inefficient. He suggested that it be meridional circulation (\\S 199). His argument for the importance of meridional circulation was qualitative. In a later paper he evaluated quantitatively the meridional circulation velocity without further commenting on its efficiency in mixing stellar interiors \\citep{Eddington29}. While it was realized in the 1940s that the presence of the giant branch in clusters implied that stars could not be completely mixed, Eddington's argument seems to have had enough weight to prevent the proper calculation of the effect of differential radiation pressure until the late 1960s even if Eddington had partially corrected his argument (see the 1930 correction on page xiii of \\citealt{Eddington26}). Work however continued on gravitational settling in outer solar layers \\citep{Biermann37,Wasiutynski58,AllerCh60}; the most important application was made to white dwarfs \\citep{Schatzman45}. Instead of assuming that equilibrium concentrations were reached, one of us introduced the differential radiative acceleration term, \\gr, into the diffusion velocity equation of \\citet{AllerCh60} and calculated anomalies to be expected in atmospheric regions \\citep{Michaud70}. Comparison to observations of ApBp stars suggested that \\gr{} plays a role in at least some stars. As large atomic data bases started becoming available in the 1980s, we realized that it would be possible to calculate \\gr{} throughout stellar interiors with good accuracy at the same time as the evolution proceeds. All composition changes can then be taken into account self consistently during evolution for the \\gr{} and the Rosseland averaged opacity. It was first tried to make those calculations with TOPBase from the Opacity Project \\citep{AlecianMiTu93,LeBlancMi95,GonzalezArMi95,GonzalezLeAretal95} but using those data we could not reproduce the concentration dependence of the Rosseland averaged OPAL opacities around the solar center (see \\S 5.1 of \\citealt{TurcotteRiMietal98}) and we shifted to using OPAL spectra \\citep{RicherMiRoetal98} to calculate \\gr{}. The original spectra they had used to calculate the OPAL opacity tables \\citep{IglesiasRo93,IglesiasRo96,RogersIg92a} were kindly made available to us by Iglesias and Rogers. In this review of \\gr{} in stellar evolution, we will first briefly describe how the calculations are carried out in stellar evolution codes (\\S\\,\\ref{sec:calculations}) and compare the values of \\gr{} obtained using OP and OPAL data (\\S\\,\\ref{sec:CorrectionFactors}). A few examples of the effect of \\gr{} in Pop I and Pop II stars are described (\\S\\,\\ref{sec:Role}) and, finally, results obtained recently for HB stars are presented (\\S\\,\\ref{sec:HBStars}) showing that \\gr{}s play a role in stellar evolution (\\S\\,\\ref{sec:con}). \\begin{figure*}[t!] \\resizebox{\\hsize}{!}{\\includegraphics[clip=true]{OPcorr_grad.eps}} \\caption{\\footnotesize Correction factors used in evolution calculations except for Fe, see the text. Adapted from Fig. 6 of \\citealt{RicherMiRoetal98}. }. \\label{fig:correction} \\end{figure*} ", "conclusions": "\\label{sec:con} The availability of large atomic data bases has allowed, as described in \\S\\,\\ref{sec:calculations}, to calculate stellar evolution models for Pop I and II stars up and past the giant branch including all effects of atomic diffusion. They cause abundance anomalies not only in Ap stars, as originally suggested by \\citet{Michaud70}, but also in HB stars of clusters (\\S \\ref{sec:SurfaceAbundanceAnomalies}) and possibly in turnoff stars (\\S \\ref{sec:Role}). They play an essential role in driving pulsations in sdB stars \\citep{FontaineBrChetal2003}, the field analogue of HB stars. It is furthermore not only the surface region which is affected but 50\\% of the stellar radius and 10$^{-3}$ of its mass \\citep{RichardMiRi2002,MichaudRiRi2007}. \\citet{Eddington26} was right however in suggesting that competing processes also have a role to play. For instance, in M15, rotation plays a role probably through meridional circulation. The arrow in Figure \\ref{fig:Fe} shows the only star with $\\teff> 11500$\\,K which has no abundance anomaly. It is also the only relatively rapidly rotating star. \\citet{QuievyChMietal2007} have shown that meridional circulation explained that the stars with $\\teff<11500$\\,K have a normal abundance and not the 5$\\times$ overabundance that Figure \\ref{fig:Fe} would suggest. While atomic diffusion driven by \\gr{}s plays the main role in creating abundance anomalies on the HB, it has to compete with the effects of rotation just as in HgMn or AmFm stars \\citep{CharbonneauMi91}." }, "0801/0801.1670_arXiv.txt": { "abstract": "We present a study of the mass-metallicity ($M-Z$) relation and H~II region physical conditions in a sample of 20 star-forming galaxies at $1.0 < z < 1.5$ drawn from the DEEP2 Galaxy Redshift Survey. We find a correlation between stellar mass and gas-phase oxygen abundance in the sample and compare it with the one observed among UV-selected $z \\sim 2$ star-forming galaxies and local objects from the Sloan Digital Sky Survey (SDSS). This comparison, based on the same empirical abundance indicator, demonstrates that the zero point of the $M-Z$ relationship evolves with redshift, in the sense that galaxies at fixed stellar mass become more metal-rich at lower redshift. Measurements of [O III]/H$\\beta$ and [N II]/H$\\alpha$ emission-line ratios show that, on average, objects in the DEEP2 $1.0 < z < 1.5$ sample are significantly offset from the excitation sequence observed in nearby H~II regions and SDSS emission-line galaxies. In order to fully understand the causes of this offset, which is also observed in $z\\sim2$ star-forming galaxies, we examine in detail the small fraction of SDSS galaxies that have similar diagnostic ratios to those of the DEEP2 sample. Some of these galaxies indicate evidence for AGN and/or shock activity, which may give rise to their unusual line ratios and contribute to Balmer emission lines at the level of $\\sim 20$\\%. Others indicate no evidence for AGN or shock excitation yet are characterized by higher electron densities and temperatures, and therefore interstellar gas pressures, than typical SDSS star-forming galaxies of similar stellar mass. These anomalous objects also have higher concentrations and star formation rate surface densities, which are directly connected to higher interstellar pressure. Higher star formation rate surface densities, interstellar pressures, and H~II region ionization parameters may also be common at high redshift. These effects must be taken into account when using strong-line indicators to understand the evolution of heavy elements in galaxies. When such effects are included, the inferred evolution of the $M-Z$ relation out to $z\\sim 2$ is more significant than previous estimates. ", "introduction": "The abundance of heavy elements in the H~II regions of galaxies reflects the past history of star formation and the effects of inflows and outflows of gas. A characterization of the evolution of chemical abundances for galaxies of different masses is therefore essential to a complete model of galaxy formation that includes the physics of baryons \\citep{delucia04,finlator07}. Important observational constraints for such models come from determining the scaling relations at different redshifts among galaxy luminosity, stellar mass, and metallicity, which, for star-forming galaxies, typically consists of the oxygen abundance. However, one of the key challenges is to take the observationally measured quantities, i.e. strong, rest-frame optical emission-line ratios, and connect them with the physical quantity of interest, i.e. oxygen abundance. \\begin{deluxetable*}{lcccccccc} \\tablewidth{0pt} \\tabletypesize{\\footnotesize} \\tablecaption{Galaxies Observed with Keck~II NIRSPEC\\label{tab:obs}} \\tablehead{ \\colhead{~~~~~~~~~~DEEP ID~~~~~~~~~~} & \\colhead{~~~~~R.A. (J2000)~~~~~} & \\colhead{~~~~~Dec. (J2000)~~~~~} & \\colhead{~~~~~$z_{{\\rm H}\\alpha}$~~~~~} & \\colhead{~~~~~$B$~~~~~} & \\colhead{~~~~~$R$~~~~~} & \\colhead{~~~~~$I$~~~~~} & \\colhead{~~~~~$M_B$~~~~~} & \\colhead{~~~~~$U-B$~~~~~} } \\startdata 42044579 \\dotfill & 02 30 43.46 & 00 42 43.60 & 1.0180 & 23.22 & 22.97 & 22.40 & -21.20 & 0.54 \\\\ 22046630 \\dotfill & 16 50 13.83 & 35 02 01.78 & 1.0225 & 23.64 & 23.02 & 22.31 & -21.37 & 0.69 \\\\ 22046748 \\dotfill & 16 50 14.55 & 35 02 04.31 & 1.0241 & 24.43 & 23.76 & 22.90 & -20.86 & 0.86 \\\\ 42044575 \\dotfill & 02 30 44.85 & 00 42 51.33 & 1.0490 & 23.08 & 22.94 & 22.56 & -21.06 & 0.34 \\\\ 42010638 \\dotfill & 02 29 08.74 & 00 23 28.40 & 1.3877 & 22.93 & 22.85 & 22.54 & -22.12 & 0.49 \\\\ 42010637 \\dotfill & 02 29 08.74 & 00 23 32.87 & 1.3882 & 24.20 & 23.98 & 23.72 & -20.87 & 0.44 \\\\ 42021412 \\dotfill & 02 30 44.55 & 00 30 50.73 & 1.3962 & 24.07 & 23.74 & 23.12 & -21.91 & 0.78 \\\\ 42021652 \\dotfill & 02 30 44.70 & 00 30 46.19 & 1.3984 & 22.97 & 22.24 & 21.32 & -24.01 & 1.01 \\\\ \\enddata \\tablecomments{Units of right ascension are hours, minutes, and seconds, and units of declination are degrees, arcminutes and arcseconds.} \\end{deluxetable*} In the local universe, \\citet{tremonti04} have used a sample of $\\sim$53,000 emission-line galaxies from the Sloan Digital Sky Survey (SDSS) to investigate the luminosity-metallicity ($L-Z$) and mass-metallicity ($M-Z$) relationships. For this sample, metallicities were estimated on the observed spectra of several strong emission lines, including [O II] $\\lambda\\lambda$3726, 3729, H$\\beta$, [O III]$ \\lambda\\lambda$5007, 4959, H$\\alpha$, [N II] $\\lambda\\lambda$6548, 6584, and [S II] $\\lambda\\lambda$6717, 6731. At increasing redshifts, as the strong rest-frame optical emission lines shift into the near-IR, metallicities are typically based on smaller subsets of strong emission lines through the use of empirically calibrated abundance indicators \\citep[e.g.][]{pp04,pagel79}. Much progress has been made recently in assembling large samples of star-forming galaxies with abundance measurements at both intermediate redshift \\citep{kewley02,savaglio05} and at $z>2$ \\citep{erb06a}. However, we have only begun to gather chemical abundance measurements for galaxies at $z \\sim 1-2$ \\citep[][hereafter Paper I,]{maier06,shapley05}. In this work, we continue our efforts to fill in the gap of chemical abundance measurements during this important redshift range, which hosts the emergence of the Hubble sequence of disk and elliptical galaxies \\citep{dickinson00}, and the buildup of a significant fraction of the stellar mass in the universe \\citep{drory05,dickinson03} prior to the decline in global star formation rate (SFR) density \\citep{madau96}. Chemical abundances for high-redshift galaxies are commonly estimated using locally calibrated empirical indicators. Yet it is crucial to recognize the fact that a considerable fraction of the $z \\sim 1$ and $2$ galaxies with measurements of multiple rest-frame optical emission lines do not follow the local excitation sequence described by nearby H~II regions and star-forming galaxies in the diagnostic diagram featuring the [O III] $\\lambda$5007/H$\\beta$ and [N II] $\\lambda$6584/H$\\alpha$ emission-line ratios \\citep[Paper I;][]{erb06a}. On average, the distant galaxies lie offset towards higher [O III] $\\lambda$5007/H$\\beta$ and [N II] $\\lambda$6584/H$\\alpha$, relative to local galaxies. As discussed in Paper I and \\citet{groves06}, several causes may account for this offset, in terms of the prevailing physical conditions in the H~II regions of high-redshift galaxies. The relevant conditions are H~II region electron density, hardness of the ionizing spectrum, ionization parameter, the effects of shock excitation, and contributions from an active galactic nucleus (AGN). It is still unclear which of these are most important for determining the emergent spectra of high- redshift galaxies. Understanding this offset in emission-line ratios is important, not only because it provides evidence that physical conditions in the high redshift universe are different from the local ones, but also because the application of an empirically calibrated abundance indicator to a set of H~II regions or star-forming galaxies rests on the assumption that these objects are similar, on average, to those on which the calibration is based. In this sense, the current work is also motivated by the interpretation of the offset in emission-line ratios among distant galaxies, and an assessment of the reliability of using local abundance calibrations for high-redshift star-forming galaxies. Instead of focusing on high-redshift objects, another approach is to study the properties in a class of nearby objects, which exhibit similar offset behavior on the emission-line diagnostic diagram. Unravelling the relations between the physical conditions and unusual diagnostic line ratios for such objects aids the understanding of high-redshift galaxies. The SDSS, with its rich set of photometric and spectroscopic information, provides an ideal local comparison sample. In this paper we expand on the analysis presented in Paper I, with an enlarged sample of DEEP2 star-forming galaxies observed with NIRSPEC on the Keck II telescope. The larger number of DEEP2 objects with near-IR observations enforces the conclusions drawn in the previous work. Furthermore, our detailed study of nearby SDSS objects with similar emission-line diagnostic ratios leads to a clearer physical explanation of the observed properties of the DEEP2 galaxies. The DEEP2 sample, near-IR spectroscopic observations, data reduction, and measurements are described in \\S 2. We present the oxygen abundances derived from measurements of [O III], H$\\beta$, H$\\alpha$, and [N II] emission lines in both individual as well as composite spectra in \\S 3. The mass-metallicity relationship and its evolution through cosmic time are discussed in \\S 4. In \\S 5 we investigate differences in $z \\sim 1.0-1.5$ H II region physical conditions with respect to local samples by examining nearby SDSS galaxies with similar emission-line diagnostic ratios. Finally, we summarize our main conclusions in \\S 6. A cosmology with $\\Omega_m = 0.3$, $\\Omega_{\\Lambda} = 0.7$, and $h = 0.7$ is assumed throughout. ", "conclusions": "We have compiled a sample of 20 star-forming galaxies at $1.0 < z < 1.5$ drawn from the blue cloud of the color bimodality observed in the DEEP2 survey, to study the correlation between stellar mass and metallicity, across a dynamical range of 2 orders of magnitude in stellar mass, as well as H~II region physical conditions at this redshift range. In order to gain some insights on the causes of the offset in the BPT diagram observed in high-redshift star-forming galaxies, we have examined the H II region diagnostic line ratios and host galaxy properties of the small fraction of SDSS galaxies that have similar diagnostic ratios to those of the DEEP2 sample. Our main results are summarized as follows: \\begin{enumerate} \\item[1.] There is a correlation between stellar mass and gas-phase oxygen abundance in DEEP2 star-forming galaxies at $z \\sim 1.0$ and at $z \\sim 1.4$. We have found that the zero point of the $M-Z$ relationship evolves with redshift, in the sense that galaxies at fixed stellar mass become more metal-rich at lower redshift, by comparing the $1.0 < z < 1.5$ sample with UV-selected $z \\sim 2$ and SDSS local star-forming galaxies. At the low-mass end ($M_{\\star} \\sim 8\\times10^9 M_{\\odot}$), the relation at $1.0 < z < 1.5$ is offset by $\\sim$0.2 (0.35) dex from the local mass-metallicity relation according to the N2 (O3N2) indicator. The N2-based offset could be larger by as much as $\\sim$0.16 dex, when the systematic bias due to difference in H II region physical conditions between $1.0 < z < 1.5$ and the local universe is taken into account. At the high-mass end ($M_{\\star} \\sim 5\\times10^{10} M_{\\odot}$), the metallicity offset between the DEEP2 $1.0 < z < 1.5$ sample and the local SDSS sample is at least $\\sim 0.2$ dex, according to the O3N2 indicator. \\item[2.] As observed previously for a very small sample of high-redshift galaxies, on average our new DEEP2 sample at $1.0 < z < 1.5$ is offset from the excitation sequence formed by nearby H~II regions and SDSS emission-line galaxies. By examining the small fraction of SDSS galaxies that have similar diagnostic ratios to those of the DEEP2 sample, we have found two likely causes for the anomalous emission-line ratios. One is the contribution from AGN and/or shock excitation at the level of $\\sim 20$\\%. The other is the difference in H~II region physical conditions, characterized by significantly larger ionization parameters, as a result of higher electron densities and temperatures, and hence higher interstellar ambient pressure, than the typical values of local star-forming galaxies with similar stellar mass. These unusual physical conditions are possibly connected to the host-galaxy properties, most importantly smaller sizes and higher star-formation rate surface densities. Our conclusion drawn from analyzing the SDSS data has been further verified by the fact that the DEEP2 objects more offset from the local excitation sequence in the BPT diagram also have higher electron densities than those closer to the local sequence. We cannot rule out either the contribution from AGN and/or shock excitation, or the difference in H~II region physical conditions, for the unusual emission-line diagnostic ratios of high-redshift star-forming galaxies. \\item[3.] We have quantified the effects of different H II region physical conditions on the strong-line metallicity calibrations. The direct electron temperature method was used to estimate the ``true'' metallicity difference between offset SDSS objects with anomalous line ratios and more typical objects of similar stellar mass. Strong-line indicators were also used to estimate this difference. A comparison of these results reveals potential biases in the strong-line indicators. According to our test, the metallicities based on strong-line indicators can either be systematically too large by $\\sim$0.16 dex (N2, N2O2), or too low by $\\sim$0.12 dex ($R_{23}$), for objects with similar H~II region physical conditions to those observed in high-redshift galaxies. The bias of O3N2 is much less significant (too large by $\\sim$0.02 dex) for offset SDSS objects with anomalous line ratios, although it may not always be negligible given other ranges of physical conditions in terms of metallicity, ionization parameter, and electron density. \\end{enumerate} The difference in H~II region physical conditions, which may commonly apply to $z\\sim 1.0-1.5$ star-forming galaxies, must be taken into account when strong-line abundance indicators are used to study the evolution of galaxy metallicity with redshift. There are at least two methods to remove the systematic bias from the effect of significantly different H~II region physical conditions on the strong-line abundance calibrations. One is to gather the abundance information with direct $T_e$ method for a sample of high-redshift H~II regions as the calibration sample, which may be hard to achieve, as auroral lines are difficult to measure except for very metal-poor objects or in deep, composite spectra. The other, which is currently feasible yet relies on photoionization models, is to quantify the biases in strong-line indicators when certain physical conditions are present and then compensate the biases when inferring abundances from strong-line indicators for high-redshift galaxies. In this study we have presented evidence that high-redshift star-forming galaxies possess distinct H~II region physical properties, as characterized by on average larger ionization parameters, higher electron densities, and temperatures, which are possibly connected to their relatively smaller sizes and higher SFR surface densities. These conditions may be quite common during the epoch at $z \\geq 1$ when at least 50\\% of the local stellar mass density was formed \\citep{bundy06,drory05}. Therefore, they should be characterized in more detail for a full understanding of the star formation history of the universe as well as the buildup of heavy elements in galaxies. The next generation of ground-based near-IR multi-object spectrographs will play a key role in assembling rest-frame optical emission-line measurements for large samples of high-redshift galaxies, enabling the detailed study of star-forming galaxies in the early universe." }, "0801/0801.4749_arXiv.txt": { "abstract": "We present a systematic survey of the temporal and spectral properties of all GRB X-ray afterglows observed by Swift-XRT between January 2005 and July 2007. We have constructed a catalog of all light curves and spectra and investigate the physical origin of each afterglow segment in the framework of the forward shock models by comparing the data with the closure relations. We search for possible jet-like breaks in the lightcurves and try to explain some of the \"missing\" X-ray jet breaks in the lightcurves. ", "introduction": "Studies of the presence or absence of jet breaks in GRB X-ray afterglows have been recently undertaken with several different approaches yielding differing results \\citep{burrows07,liang07b,panaitescu07,kocevski07}. The importance of this work lies in that the results have vital implications on the energetics, geometry, and frequency of GRBs. The fact that they do not behave as expected from pre-{\\it Swift} observations is not surprising in the context of how much we do not understand and have only recently learned about all of the aspects of X-ray afterglows. To understand the jet break phenomena we must understand it in the global context of GRB and afterglow properties. Therefore, the goal of our study is to do a census of X-ray afterglow properties by fitting a variety of physical models to each component of the afterglows and understand how the jet breaks fit in as one component in the larger coherent picture of this phenomenon. ", "conclusions": "Although the jet break phenomena is not fully understood, we are beginning to be able to better explain the paucity of expected observations of them. We are able to identify a sizeable group of afterglows that likely contain jet breaks or post-jet break data even though they do not present themselves in the classical context of the full canonical model. Due to observational limitations and additional possible physical model variations, even more afterglows may be consistent with the expectations from those that do contain confident jet breaks, but are currently indistinguishable. While we are beginning to understand or at least be able to explain the majority of our sample, we are also finding an interesting small subset of outliers that confidently do not contain jet breaks during the time interval in which we would expect to see them. These afterglows require further investigation and perhaps are somehow fundamentally different in their jet and afterglow properties. \\begin{theacknowledgments} JLR, DNB, DCM, and AF acknowledge support under NASA contract NAS5-00136. \\end{theacknowledgments}" }, "0801/0801.2390_arXiv.txt": { "abstract": "We present {\\em GALEX} near-UV ($NUV$) and 2MASS $J$ band photometry for red sequence galaxies in local clusters. We define quiescent samples according to a strict emission threshold, removing galaxies with very recent star formation. We analyse the $NUV$--$J$ colour--magnitude relation (CMR) and find that the intrinsic scatter is an order of magnitude larger than for the analogous optical CMR ($\\sim$0.35 rather than 0.05 mag), in agreement with previous studies. Comparing the $NUV$--$J$ colours with spectroscopically-derived stellar population parameters, we find a strong ($> 5.5 \\sigma$) correlation with metallicity, only a marginal trend with age, and no correlation with the $\\alpha$/Fe ratio. We explore the origin of the large scatter and conclude that neither aperture effects nor the UV upturn phenomenon contribute significantly. We show that the scatter could be attributed to simple `frosting' by either a young or a low metallicity subpopulation. ", "introduction": "\\label{sec:intro} The optical colour--magnitude relation (CMR) shows that brighter early-type galaxies are also redder \\citep{vis77}, and is traditionally regarded as arising from the mass--metallicity sequence (cf. \\citealt{dre84}, \\citealt{kod97}). Gas loss, caused by supernova wind, occurs earlier in less massive galaxies. Therefore, a smaller fraction of gas is processed before being expelled from the less massive galaxy \\citep{mat71}, resulting in lower average metallicities \\citep{lar74}. \\citet{bow92} found a very small intrinsic scatter in the $U$--$V$ CMR ($\\sim$0.05 mag) and, due to the sensitivity of the $U$ band to the presence of young stars, interpreted this as a small age dispersion. Age and metallicity are observed to have the same effect on broadband optical colours, whereas spectral line indices can be used to break the degeneracy \\citep{wor94}. \\citet{kun98} claimed that the CMR is driven by metallicity variations with luminosity, although \\citet{nel05} found evidence for a strong age--mass relation in addition to this metallicity--mass trend (see also \\citealt{cal03}, \\citealt{tho05}). The ultraviolet--optical CMR for non-star-forming galaxies has an intrinsic scatter an order of magnitude larger than its optical counterpart; $\\sim$0.5 mag compared to 0.05 mag (e.g. \\citealt{yi05}). Hot young stellar populations dominate the ultraviolet (UV) flux for $\\sim$100 Myr after an episode of star formation (ten times longer than H$\\alpha$ emission after star formation; \\citealt{lei99}). The large intrinsic scatter in the UV CMR is therefore often interpreted as differing quantities of very recent, albeit low level, star formation \\citep{fer00}. In intermediate age populations ($\\sim$1--3 Gyr), the near-UV (NUV; 2000--3000\\AA) flux is dominated by hot stars on the main sequence turn-off \\citep[e.g.][]{oco99}. The sensitivity of the turn-off to the epoch of formation emphasises the importance of the UV bands for age determination \\citep{dor03}. Old ($\\sim$10 Gyr) metal-poor populations have a significant UV flux contribution from very hot ($T_{\\rm eff}$ $\\sim$ 10000K) blue horizontal branch (BHB) stars \\citep{mara00, lee02}. However these tend to reside in globular clusters or galactic haloes (where Fe/H $< -1$), where they are useful age indicators \\citep{kav07}, rather than in relatively metal-rich elliptical galaxies. The UV picture is further complicated by the presence of the ultraviolet upturn (or UV excess, UVX) phenomenon. First observed by \\citet{cod69}, this unanticipated upturn dominates the far-UV (FUV; $<$ 2000\\AA) in UVX galaxies. In contrast, the NUV can be decomposed into two separate components: the blue end of the main sequence/subgiant branch, and the UVX contribution \\citep{dor97}. \\citet{bur88} found that the UVX can sometimes be appreciable at wavelengths as long as 2700\\AA: for example, in NGC 4649 $\\sim$75 per cent of the NUV flux can be attributed to the UVX component. However, the UVX cannot be explained by the BHB population, as the temperature required to fit the upturn would be $T_{\\rm eff}$ $\\ga$ 20000K, whereas BHBs are usually no hotter than $T_{\\rm eff}$ $\\sim$ 12000K \\citep{oco99}. \\citeauthor{bur88} further reported that FUV flux (assumed to trace the UVX) is strongly correlated with the Mg$_2$ line strength ($\\sim$metallicity) and also with the velocity dispersion, which is a proxy for galaxy mass. However, more recent studies (e.g. \\citealt{ric05}) have weakened the case for a strong UVX vs metallicity relation. From analysis of internal colour gradients, \\citet{oco92} concluded that the FUV flux in most early-types originates from old stellar components. Drawing on these results, the source of the UVX is tentatively identified as hot, low mass, helium burning stars, such as extreme horizontal branch (EHB) or `failed' AGB (AGB-manqu\\'{e}) stars and their progeny (see \\citealt{yi97}, or the review \\citealt{oco99}). The {\\em Galaxy Evolution Explorer} ({\\em GALEX}; launched in 2003; \\citealt{mar05}, \\citealt{mor07}) is revolutionising UV astronomy, with high resolution imaging in two bands: near-ultraviolet ($NUV$; $\\lambda_{eff} = 2310$ \\AA) and far-ultraviolet ($FUV$; $\\lambda_{eff} = 1530$ \\AA). Using analysis of both $NUV$--$V$ and $FUV$--$V$ vs $B$--$V$ relations, \\citet{don06} suggest that the $FUV$--$NUV$ colour reflects an extension of the colour--metallicity relation into the UV, as well as deducing that $\\sim$10 per cent of ellipticals have residual star formation. Using the $NUV$--$\\,r$ colour, \\citet{kav06} also find non-negligible young stellar populations in morphologically selected early-type galaxies. \\citet{sal07} derive star formation rates (SFRs) from broadband photometry dominated by the UV, and find good agreement with SFRs deduced from spectroscopic indices (predominantly using H$\\alpha$). However, they also confirm that some galaxies with no H$\\alpha$ emission show signs of star formation in the UV bands and attribute this to post-starburst galaxies. Here, we build upon these previous studies by exploring the relationship between the $NUV$--$J$ colour and spectroscopic stellar population indicators for a sample of quiescent, red sequence galaxies in nearby clusters. This paper is organised as follows. Section \\ref{sec:data} describes our two red sequence samples and associated 2MASS and {\\em GALEX} datasets. The criteria used to remove galaxies with emission are described. In Section \\ref{sec:results} we show that a large intrinsic scatter is found in the $NUV$--$J$ colours of these quiescent cluster galaxies. Metallicity is shown to be strongly correlated with the $NUV$--$J$ colour, although there remains a large residual scatter. Section \\ref{sec:discuss} discusses possible explanations for this scatter, showing that morphological abnormalities, aperture bias and the UV upturn do not contribute significantly. We investigate simple `frosting' models with a low mass fraction of younger stars (or alternatively a low mass fraction population of low metallicity, blue horizontal branch stars), and show that these could account for the scatter. The uncertainties in the $NUV$ K-correction are also discussed. Our conclusions are given in Section \\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} Using {\\em GALEX} UV and 2MASS $J$ band photometry, we have investigated the relationship between UV--IR colours and spectroscopically derived stellar population parameters (age, metallicity and $\\alpha$-abundance) for red sequence galaxies in local clusters. We select galaxies using strict emission criteria to avoid contamination from galaxies with very recent star formation. We analyse the $NUV$ colour--magnitude relation (CMR) for our two samples of quiescent galaxies (920 in NFPS; 156 in SSC), and find rms dispersions of 0.37 and 0.30 mag (intrinsic scatter of 0.36 and 0.29 mag) respectively. This is similar to previously reported values of $\\sim$0.5 mag and is an order of magnitude larger than the scatter in the optical CMR ($\\sim$0.05 mag). We compared the $NUV$--$J$ colour to the spectroscopic stellar population parameters for 87 galaxies in the SSC sample and found the following: \\begin{itemize} \\item There is a significant $NUV$--$J$ vs metallicity trend, with a slope of $1.27 \\pm 0.23$ and an rms dispersion of 0.32 mag. \\item There is only a weak $NUV$--$J$ vs age trend after the metallicity effect has been removed, and no correlation with $\\alpha$-abundance. \\item There is a large intrinsic scatter ($\\sim$0.25 mag) in the $NUV$--$J$ colour at fixed age and metallicity which cannot be easily accounted for with simple stellar populations. \\end{itemize} The unexpected blue colours of at least some objects, including an influential outlier, cannot be attributed to large radius contamination from other objects, and aperture bias cannot account for the large scatter. Corrections for the UV upturn (UVX) phenomenon are relatively small ($\\sim$0.2 mag) and are similar galaxy-to-galaxy, so do not reduce the intrinsic scatter. We find that the large $NUV$--$J$ intrinsic scatter could be attributed to galaxy `frosting' by small ($<$ 5 per cent) populations of either young stars or a low metallicity blue horizontal branch." }, "0801/0801.0489_arXiv.txt": { "abstract": "{There is ${}^3P_2$ neutron superfluid region in NS (neutron star) interior. For a rotating NS, the ${}^3P_2$ superfluid region is like a system of rotating magnetic dipoles. It will give out electromagnetic radiation, which may provides a new heating mechanism of NSs. This heating mechanism plus some cooling agent may give sound explanation to NS glitches. ", "introduction": "Since the discovery of NSs (neutron star) and glitches in late 1960s, it is generally believed that there is superfluidity in NS interior (Ruderman 1976, Shapiro et al. 1983, Elgar{\\o}y et al. 1996). Cooling mechanism associated with superfluidity was first proposed by Flowers et al. (1976). Not until recently, the importance of superfluidity is considered seriously in the \"Minimal model\" (Gusakov et al. 2004, Page et al. 2004, Kaminker et al. 2006). Then one may ask: since cooling agent associated with superfluidity must be considered in the \"Minimal model\", what about the heating mechanisms? Heating mechanism accompanied with superfluidity has been discussed by Alpar et al. (1989), and was taken into NS cooling model by Umeda et al (1994), Page et al. (2006), Tsuruta (2006). Here we investigate another possibility: microscope magnetic dipole radiation (MMDR) in NSs, possibly another heating mechanism associated with superfluidity. The idea is as follows: there is ${}^3P_2$ superfluid region in the interior of a NS. We calculate its paramagnetic properties in the presence of a background magnetic field. Since the neutron superfluid is in a vortex state, the ${}^3P_2$ superfluid region is like a system of rotating magnetic dipoles. This system will give magnetic dipole radiation. The emitted photon can not penetrate the NS matter, thus provides a heating mechanism of NSs. The origin of this heating mechanism is microscopic magnetic dipole radiation, so we call it MMDR heating of NSs. This heating mechanism plus some cooling agent may give sound explanation to NS glitches (Bai et al. 2006; Peng et al. 2006; Peng 2007). Before going into detailed calculations of MMDR, we will first look at the superfluidity in NSs. ", "conclusions": "Here we have presented another possible heating mechanism of NSs, associated with superfluidity. It can be compared with other heating mechanisms and cooling agents of NSs (Gusakov et al. 2004, Page et al. 2006). \\begin{figure}[!htbp] \\centering \\includegraphics[width=\\textwidth]{MMDRHeating.eps}\\\\ \\caption{MMDR Heating VS Cooling. The brown one represents MMDR heating. It increase with decreasing temperature, as a result of increasing thermal factor. The red, green, and black one corresponds to photon cooling, MUrca (Modified Urca) process, and PBF (Pair Breaking and Formation) process respectively. (Adapted from Gusakov et al. (2004).)} \\label{MMDR} \\end{figure} As shown in Fig. \\ref{MMDR}, MMDR heating may be dominate only in the photon cooling stage. So it will not affect the cooling scenario of young and mid age NSs. But for old NSs, e.g. PSR 1055-52, it may serve as a moderate heating (Tsurata 2006, Page 2006). Using toy model given by Yakovlev et al. (2003), we can make illustrative calculation of NS cooling, including the MMDR heating. \\begin{figure}[!htbp] \\centering \\includegraphics[width=\\textwidth]{CoolingCurves.eps}\\\\ \\caption{Cooling Curves Including MMDR Heating. The lower and upper curve correspond to MUrca and bremsstrahlung process dominated case respectively. Whereas the plateau at latter stage is due to MMDR heating. The physical input are the same as Yakovlev et al. (2003) except that we include the MMDR heating. Remind: these cooling curves are for illustrative use only.} \\label{CoolingCurves} \\end{figure} Fig. \\ref{CoolingCurves} shows that, there is a plateau in the late stage of NS cooling curves. This plateau can be compared with that of Kang's (2007). This MMDR heating must cease after some time. In our case, the cease of MMDR heating has several possibilities. \\begin{enumerate} \\item As shown by Huang et al. (1982), the energy input of MMDR heating is through Ekman pumping. When this agent is out of work, MMDR heating have to make a pause. \\item When there is a phase transition in the core, e.g. deconfinement of hardrons, all the hadron processes disappear including MMDR heating. \\item Due complexities of superfluidity gap (Elagr{\\o}y et al. 1996), the ${}^3P_2$ superfluidity region becomes slimer when the core becomes more compact. The MMDR heating contribution is ignorable in this case. \\end{enumerate} We presented another possible NS heating mechanism associated with superfluidity. The exact effect of MMDR heating needs accurate calculation of the cooling curves. This heating mechanism plus some cooling agent may give sound explanation to NS glitches. The detailed investigation is the scope of another work (Peng 2007)." }, "0801/0801.0440_arXiv.txt": { "abstract": "We compare the stellar wind torque calculated in a previous work (Paper~II) to the spin-up and spin-down torques expected to arise from the magnetic interaction between a slowly rotating ($\\sim 10$\\% of breakup) pre-main-sequence star and its accretion disk. This analysis demonstrates that stellar winds can carry off orders of magnitude more angular momentum than can be transferred to the disk, provided that the mass outflow rates are greater than the solar wind. Thus, the equilibrium spin state is simply characterized by a balance between the angular momentum deposited by accretion and that extracted by a stellar wind. We derive a semi-analytic formula for predicting the equilibrium spin rate as a function only of the ratio of $\\dot M_{\\rm w} / \\dot M_{\\rm a}$ and a dimensionless magnetization parameter, $\\Psi \\equiv B_*^2 R_*^2 (\\dot M_{\\rm a} v_{\\rm esc})^{-1}$, where $\\dot M_{\\rm w}$ is the stellar wind mass outflow rate, $\\dot M_{\\rm a}$ the accretion rate, $B_*$ the stellar surface magnetic field strength, $R_*$ the stellar radius, and $v_{\\rm esc}$ the surface escape speed. For parameters typical of accreting pre-main-sequence stars, this explains spin rates of $\\sim 10$\\% of breakup speed for $\\dot M_{\\rm w} / \\dot M_{\\rm a} \\sim 0.1$. Finally, the assumption that the stellar wind is driven by a fraction of the accretion power leads to an upper limit to the mass flow ratio of $\\dot M_{\\rm w} / \\dot M_{\\rm a} \\la 0.6$. ", "introduction": "\\label{sec_intro} The slow rotation rates of low to intermediate mass ($\\la 2 M_\\odot$) pre-main-sequence stars remains one of the most important aspects of star formation that has, so far, resisted a generally accepted explanation. By the time they become optically visible as T Tauri stars \\citep[TTSs;][]{joy45}, approximately half of them are observed to rotate at approximately 10\\% of breakup speed \\citep[the ``slow rotators''; e.g.,][]{vogelkuhi81, bouvier3ea97, rebull3ea04, herbstea07}. This is a surprise because many TTSs (the Classical T Tauri stars; CTTSs) are actively accreting material from surrounding Keplerian disks \\citep{lyndenbellpringle74, bertout3ea88, calvetgullbring98, muzerolle3ea01}. At a typical accretion rate of $\\dot M_{\\rm a} \\sim 10^{-8} M_\\odot$ yr$^{-1}$, the angular momentum deposited by accreting disk material should spin up a CTTS to near breakup speed in $\\sim 10^6$ years \\citep{hartmannstauffer89, mattpudritz07coolstars}. Since the accretion phase lasts for $10^6$ -- $10^7$ years \\citep{lyolawson05, jayawardhanaea06}, since the stars accrete at much higher rates prior to the TTS phase, and since the stars are still contracting \\citep{rebullea02}, an efficient angular momentum loss mechanism is required to explain the existence of the slow rotators. A few interesting and important ideas for explaining the TTS slow rotators have been developed over the last two decades. These have resulted in the star-disk interaction model of \\citet{ghoshlamb78}, applied to CTTSs by \\citet[][and see \\citealp{camenzind90}]{konigl91}, the X-wind model \\citep{shuea94}, and the idea that stellar winds provide strong torques \\citep{hartmannstauffer89, toutpringle92, paatzcamenzind96, ferreira3ea00, mattpudritz05l}. Although both have advanced our understanding of the magnetic star-disk interaction, neither the Ghosh \\& Lamb nor X-wind models are without problems \\citep{ferreira3ea00, uzdensky04, mattpudritz05}, and the idea that stellar winds are important has not yet been worked out in sufficient detail to compare to the other models. In \\citet[][hereafter Paper~I]{mattpudritz05l}, we further explored powerful stellar winds as a solution to the angular momentum problem and suggested that a fraction of the accretion power provides the energy necessary to drive the wind. We showed that stellar winds are capable of carrying off the accreted angular momentum, provided that $\\dot M_{\\rm w} / \\dot M_{\\rm a} \\sim 0.1$, where $\\dot M_{\\rm w}$ is the outflow rate of material that is magnetically connected to the star (the ``stellar wind''). This analysis included a formulation for the stellar wind torque that contained the Alfv\\'en radius ($r_{\\rm A}$), which is not easily determined a priori in the wind, and the conclusions were based on a one-dimensional scaling estimate of this important physical quantity. Thus, while it is clear that accretion-powered stellar winds (APSWs) can in principle provide the necessary spin-down torque, this idea requires further development to produce a more detailed model. Toward this goal, \\citet[][hereafter Paper~II]{mattpudritz08II} used 2-dimensional (axisymmetric) magnetohydrodynamic simulations to solve for $r_{\\rm A}$ and calculate realistic stellar wind torques for a range of parameters. In the present paper, we use the stellar wind solutions of Paper~II to compare the stellar wind torque to the torques expected to arise from the star-disk interaction. Furthermore, we find new solutions for stellar spins, based upon torque balance between the accretion torque and the APSW spin-down torque. This paper begins with a brief description of the simulation results of Paper~II (\\S \\ref{sec_simulations}). We then compare the stellar wind torque to the star-disk spin-down torque in section \\ref{sec_tdsd} and then to the star-disk spin-up torque in section \\ref{sec_equilibrium}, which contains spin-equilibrium solutions. Section \\ref{sec_discussion} contains a summary and discussion. ", "conclusions": "\\label{sec_discussion} In this work, we have further developed the accretion-powered stellar wind model proposed in Paper~I, where the stellar wind magnetic lever arm length, $r_{\\rm A}$, was taken as a parameter to determine the stellar wind torque. We employed the simulation results of Paper~II (see \\S \\ref{sec_simulations}) to obtain stellar wind torques for several cases representative of T Tauri systems. We examined the total torque on the star arising from the stellar wind plus the magnetic interaction between the star and its accretion disk. Our results can be summarized as follows. \\begin{enumerate} \\item{We found that the spin-down torque from a stellar wind can be orders of magnitude stronger than the spin-down portion of the star-disk interaction torque, for slowly rotating stars with mass loss rates substantially larger than the solar wind outflow rate (see \\S \\ref{sec_tdsd}). This confirms the assumption of Paper~I that the condition for net zero torque on the star (spin equilibrium) is simply determined by a balance between the stellar wind torque and the accretion torque.} \\item{Using the computed stellar wind torques for several cases, we looked at the conditions for spin equilibrium (\\S \\ref{sec_balancing}). We found that a rotation rate of 10\\% of breakup speed typically requires $\\dot M_{\\rm w} / \\dot M_{\\rm a}$ equal a few tens of percent, confirming the original suggestion by \\citet{hartmannstauffer89} that stellar winds may be capable of removing accreted angular momentum.} \\item{For most cases in spin equilibrium, the disk truncation radius was very close to the corotation radius, though this is not a general requirement of the APSW model (\\S \\ref{sec_state}).} \\item{Accretion power is generally sufficient to power a stellar wind that is capable of solving the angular momentum problem (\\S \\ref{sec_power}), as suggested in Paper~I. The energy requirements for most of the cases considered here is relatively large, and more work is needed to further constrain the energy coupling.} \\item{Under the assumption that the stellar wind is accretion powered, the cases we examined suggested a hard upper limit of $\\dot M_{\\rm w} / \\dot M_{\\rm a} \\la 0.6$.} \\item{Finally, in section \\ref{sec_semianalytic} we used the results from Paper~II to derive a semi-analytic formulation for the equilibrium spin rate predicted by the APSW model. We found the that spin rate, expressed as a fraction of breakup speed, generally depends only on the two dimensionless parameters $\\dot M_{\\rm w} / \\dot M_{\\rm a}$ and $\\Psi \\equiv B_*^2 R_*^2 (\\dot M_{\\rm a} v_{\\rm esc})^{-1}$.} \\end{enumerate} The APSW model incorporates several previous ideas. As in all other models that emphasize the role of stellar magnetic fields, the interaction of the magnetized star with the disk leads to the truncation of the disk and accretion of material along field lines onto the star. The APSW model adopts the finding of the Ghosh \\& Lamb-type models \\citep[e.g.,][]{ghoshlamb78, konigl91, armitageclarke96}, which is also supported by numerical simulations \\citep[][]{romanovaea02, long3ea05}, that the angular momentum of accreting material is transferred to the star. However, in contrast to the Ghosh \\& Lamb-type models, we found that for slow rotators, any spin-down torque arising from the star-disk interaction is negligible (item 1 above). Instead, the stellar spin-up torque from accretion is counteracted by an accretion-driven stellar wind, which carries a comparable amount of angular momentum out of the system. Compared to all existing models, APSW is distinct in that it conceptually links the driving of the stellar wind to the energy released by the accretion process (via $\\epsilon$). In other ways, the general picture of APSW is similar to other angular momentum models that utilize winds. In particular, the X-wind \\citep{shuea94, ostrikershu95}, the Reconnection X-wind \\citep{ferreira3ea00}, and other works considering stellar winds \\citep{hartmannstauffer89, paatzcamenzind96} all find that, in order to carry away significant angular momentum, the mass outflow rate needs to be of the order of 10\\% of the accretion rate (item 2 above). Except for the X-wind, in all of the above scenarios, the outflow is magnetically connected to the star, and thus extracts angular momentum directly from the star. By contrast, the X-wind outflow is magnetically connected to the disk. Furthermore, the X-wind is unique in that it assumes that the accretion of material does not deposit angular momentum onto the star. As mentioned in item 3 above, for most of the specific cases of our simulated winds, we found that in the spin equilibrium state, the truncation radius was very close to the corotation radius. This is similar to the prediction of the disk locking models, which includes both the Ghosh \\& Lamb-type models and the X-wind. However, in contrast with the disk locking models, it is not a requirement of APSW that $R_{\\rm t}$ be close to $R_{\\rm co}$. Measurements of the location of the inner edge of the gas disk \\citep{najita3ea03, carr07} suggest that $R_{\\rm t}$/$R_{\\rm co}$ is typically $\\sim 70$\\%. We leave a more detailed comparison between models for future work. There is observational evidence that the outflow rates from accreting young stellar systems are of the order of 10\\% of the accretion rates \\citep[e.g.,][]{hartigan3ea95, calvet97} and are therefore accretion powered \\citep{cabritea90}. However, it appears that a large fraction of this flow (which is usually probed by forbidden emission coming from large spatial scales) originates in the disk, rather than the star \\citep{ferreira3ea06}. It is not yet clear how much of the total observed flow may originate in a stellar wind. There is some evidence specifically for stellar winds from CTTSs \\citep{beristain3ea01, edwardsea03, dupreeea05, edwardsea06, kwan3ea07}, as distinct from disk winds, and that these are accretion powered \\citep[e.g.,][]{edwardsea03, edwardsea06}, but the mass outflow rates are not yet well constrained \\citep{dupreeea05}. Additional work constraining the value of $\\dot M_{\\rm w} / \\dot M_{\\rm a}$, the stellar wind driving mechanism, and the stellar magnetic field strength and geometry will help to provide stringent and quantitative tests for the APSW model. The predictions of the spin equilibrium state can also be checked observationally \\citep{johnskrullgafford02}. This will likely require large samples of stars, due to large uncertainties in measured parameters, and since intrinsic variability in real systems \\citep[e.g.,]{hartmann97} may only allow a spin equilibrium state to be achieved in a time-averaged sense. The Ghosh \\& Lamb, X-wind, and APSW models all predict an equilibrium spin rate that depends on $\\Psi$, but of these three, only the APSW model contains an additional dependence on the stellar wind mass outflow rate. For the power law fits to the simulations of Paper~II, the APSW spin equilibrium predicts a slightly different power-law of spin vs.\\ $\\Psi$ than the other models (\\S \\ref{sec_semianalytic})---though the exact dependence of the stellar wind torque has not been determined for all parameters. In this series of papers, we have focused on the global problem of calculating the magnitude of stellar wind torques and comparing them with other torques acting on accreting stars. In order to refine the APSW model further and make the predictions more precise, more work is required. In particular, it is not yet clear how the presence of an accretion disk will influence the stellar wind torque, and conversely, how a stellar wind may influence the accretion process. Also, although it is clear that there is enough accretion energy to power the stellar wind, it is still not known what actually drives the stellar wind and how the accretion power may transfer to it. We suspect that a strong flux of hydromagnetic waves can be excited near the base of the accretion shock and can tap the energy released there, which may provide an efficient driver for the APSW. We defer a rigourous investigation of the APSW driving mechanism to future work." }, "0801/0801.0506_arXiv.txt": { "abstract": "{} {Our goal is to demonstrate the potential of the interferometric AMBER instrument linked with the Very Large Telescope Interferometer (VLTI) fringe-tracking facility FINITO to derive high-precision stellar diameters.} {We use commissioning data obtained on the bright single star V3879~Sgr. Locking the interferometric fringes with FINITO allows us to record very low contrast fringes on the AMBER camera. By fitting the amplitude of these fringes, we measure the diameter of the target in three directions simultaneously with an accuracy of 25 micro-arcseconds.} {We showed that V3879~Sgr has a round photosphere down to a sub-percent level. We quickly reached this level of accuracy because the technique used is independent from absolute calibration (at least for baselines that fully span the visibility null). We briefly discuss the potential biases found at this level of precision.} {The proposed AMBER+FINITO instrumental setup opens several perspectives for the VLTI in the field of stellar astrophysics, like measuring with high accuracy the oblateness of fast rotating stars or detecting atmospheric starspots.} ", "introduction": "In optical, long-baseline interferometry, the most popular observable is the contrast (also called the visibility) of the interferometric fringes that appear when superposing the beams coming from several distant telescopes. The major limitation is the random optical delay introduced by the atmospheric turbulence, which make these fringes jitter around on the detector by a quantity larger than the fringe spacing. Practically, fringe blurring is avoided, either by reducing the integration time to a few milliseconds, which prevents observations of faint targets; or by using a dedicated fringe-tracking facility, the purpose of which is to stabilize fringes by measuring and correcting the optical delay in real-time. Fringe-tracking is generally used to observe fainter objects or to increase the spectral resolution, but this is not the only application. Another important gain is the possibility of recording very low-contrast fringes that that rise above the noise level after integration times longer than a few seconds. Such small fringes are produced by astronomical objects spatially resolved by the interferometer. More precisely when observing a single star modeled by a Uniform Disk of diameter \\dUD{}, the fringe visibility \\vis{} drops according to the following law: \\begin{equation} \\label{eq:0} \\visSquare{} = \\tfSquare(\\wave,t) \\;.\\;\\airySquare(\\dUD{}\\,\\blambda{}) \\end{equation} where\\begin{description} \\item[\\blambda{}] is the spatial frequency of the observation, given by the ratio of the distance between the telescopes \\base{} projected on the sky (called the baseline) and the observing wavelength \\wave{}. \\item[\\airy{}] is the Airy function (the Fourier transform of a disk of unitary size). The function \\airySquare{} actually goes to zero (fringes disappear) for a disk of diameter $\\dUD{}\\approx{}1.22/\\blambda{}$. After this first null the fringes reappear, but at lower visibility (second lobe). \\item[\\tfSquare{}] is the interferometric response of the instrumental chain, also called the transfer-function. It changes with the instrumental setup and the atmospheric conditions. Therefore, it has to be frequently calibrated by observing unresolved stars, or stars with known diameters. Statistical errors and fluctuations of this transfer-function generally dominate the error budget when trying to estimate the diameter with high accuracy. \\end{description} Because of the structure of Eq.~\\ref{eq:0}, collecting visibility points in the vicinity of the first null of the \\airySquare{} function makes the diameter estimation much less sensitive to the transfer-function uncertainty. For instance, measuring the exact spatial frequency at which the fringes disappear gives an estimation of \\dUD{}, formally independent from \\tfSquare{}, since one can use the relation: $\\dUD{}\\approx{}1.22/\\blambda{}$. In this paper, we demonstrate the feasibility of this technique at the Very Large Telescope Interferometer \\citep[VLTI, see][]{Scholler-2006jul} by using an adequate setup of the fringe-tracker FINITO \\citep{Gai-2004oct} and of the scientific instrument AMBER \\citep{Petrov-2007mar}. Section~\\ref{sec:observationsAndDataReduction} describes the instrumental setup, the observations, and the data reduction. Section~\\ref{sec:resultsAndDiscussion} details the results and discusses the obtained accuracy. The paper ends with brief conclusions and perspectives. \\begin{figure} % \\begin{flushright} \\includegraphics[scale=1]{LeBouquin_fig1.ps} \\end{flushright} \\centering \\caption{\\label{fig:technique} VLTI array configuration during the observation (top) and corresponding simulated visibility-curves (solid line, bottom), assuming the uniform disk diameter of \\vSgr{} ($7.56$~mas). FINITO tracks the fringes along D0-H0 and G1-D0 in the H-band (star symbols), while AMBER records data of the three baselines across the K-band (circle symbols). The horizontal axes are the spatial frequencies \\blambda{}, marked in meter per micron, the vertical axes are the squared and linear visibilities.} \\end{figure} ", "conclusions": "From the instrumental point of view, we have demonstrated that: \\begin{itemize} \\item AMBER is able to record and process very low-contrast fringes when they are decently locked (down to about $1.5$\\% contrast on a bright $\\mathrm{K}=\\-0.5^{\\rm m}$ star, when integrating 60 frames of 1s each); \\item FINITO is able to lock fringes in the second visibility lobe (where fringes have less than $15$\\% contrast) with a sufficiently good efficiency. \\end{itemize} As an on-sky validation, we collected K-band interferometric fringes in the vicinity of the first visibility null of the bright star \\vSgr{}. By fitting the visibility curves, we measured a diameter of $7.56\\pm{}0.025$\\,mas in three directions simultaneously. We reached such sub-percent accuracy without spending additional time on calibrators since the technique is independent from absolute calibration (at least for baselines that fully span the visibility null). We show that, at this level of precision, several systematic error sources have to be taken into account, for example the spectral calibration of the instrument and the pupil lateral position. From the scientific point of view, this work opens several perspectives for the VLTI in the field of stellar astrophysics:\\begin{itemize} \\item By using the AMBER+FINITO setup presented in this paper, one can measure a stellar diameter in three directions simultaneously with high accuracy. This can be used to constrain quickly the oblate photosphere of fast-rotating stars. \\item By using the same setup, one can precisely measure the value of the visibility minimum between the visibility lobes. Any departure from zero, even small, proves the presence of asymmetric features on (or above) the stellar surface, such as convective or magnetic starspots (in our \\vSgr{} dataset, this minimum is compatible with zero, see Fig.~\\ref{fig:fit}). \\item Finally, by associating fringe-tracking in the second lobe with baseline bootstrapping, one can record visibility points in the third visibility lobe. There, observables become very sensitive to the limb-darkening strength and its spectral dependency, which will allow thorough tests of stellar atmosphere models. \\end{itemize}" }, "0801/0801.0766_arXiv.txt": { "abstract": "We study the relationship between different wave phenomena associated with a coronal mass ejection (CME) observed on 05 Mar. 2000. EIT waves were observed in the images recorded by EIT at 195~{\\AA}. The white-light LASCO/C2 images show clear deflection and propagation of a kink along with the CME. Spectroscopic observations recorded by the UVCS reveals excessive line broadening in the two O~{\\sc{vi}} lines (1032 and 1037 {\\AA}). Moreover very hot lines such as Si ~{\\sc{xii}} and Mg ~{\\sc{x}} were observed. Interestingly, the EIT wave, the streamer deflection and the intensity modulation along the slit were all propagating North-East. Spatial and temporal correlations show that the streamer deflection and spectral line broadening are highly likely to be due to a CME-driven shock wave and that the EIT wave is the signature of a CME-driven shock wave in the lower corona. ", "introduction": "Coronal mass ejections (CMEs) are one of the most fascinating and intriguing forms of solar activity. They occur when solar plasma threaded with topologically complex magnetic fields are ejected out into the corona and interplanetary medium. Recent technological developments have yielded a steady increase in qualitative as well as quantitative studies of CMEs and related phenomena. However, numerous CME-associated phenomena such as the relationship between different wave features (EIT waves: Thompson et al. 1998; CME-driven shock waves: Hundhausen et al. 1987) are not understood unambiguously. CMEs with speed greater than the Alfv\\'en speed of the local plasma produce shock waves whose effects can be traced through radio type II bursts (Klassen et al. 2000) and/or emission of very hot lines (compared to the ambient coronal temperatures; Raouafi et al. 2004). These waves could also be detected through deflections in remote streamer belts with respect to the ejected CME material (see Hundhausen et al. 1987 and Sime \\& Hundhausen 1987). The high quality data of SOHO/LASCO (Brueckner et al. 1995) yielded numerous cases for streamer deflections associated with super-Alfv\\'enic CME eruptions (see Sheeley et al. 2000). These observations led to the conclusion deduced by the previous authors that these deflections were indeed a consequence of CME-driven shock waves. Spectroscopic signatures of coronal shock waves due to high speed CMEs ($>600$~km~s$^{-1}$; see Raymond et al. 2000) are line broadening (e.g., \\ion{O}{vi} 1032~{\\AA} \\& 1037~{\\AA} lines) and enhanced emissions in spectral lines that are rarely observed in the corona such as \\ion{Si}{xii} 520~{\\AA} and \\ion{Mg}{x} 625~{\\AA} (e.g., Raouafi et al. 2004) due to their relatively high formation temperatures ($>2$~MK). Up to now only a few CME-driven shock wave events have been reported through SOHO/UVCS (Kohl et al. 1995) observations: Raymond et al. 2000; Mancuso et al. 2002; Raouafi et al. 2004; and Ciaravella et al. 2005 (hereafter RMRC00-05). Since the discovery of the H${\\alpha}$ Moreton waves (Moreton 1964), it was thought that these waves were due to the intersection of coronal shock waves (due to flares) with the chromosphere (e.g., Uchida 1968). Later when EIT waves (Thompson et al. 1998) were discovered based on the observations recorded by EIT (Delaboudini\\`ere et al. 1995), they were interpreted as the coronal manifestation of the chromospheric Moreton wave (Thompson et al. 1999). However, the difference in propagation patterns and speeds of Moreton and EIT waves questions the similarity between these two phenomena. Although these waves are being observed more frequently the question remains open as to how these different wave phenomena are related to each other. Based on MHD simulations, (Chen et al. 2005a,b) showed that Moreton waves are the surface counter part of the CME driven shock wave. However, the EIT waves are the slow moving wave fronts traveling behind the Moreton wave due to the opening of the magnetic flux system. This was also suggested by Zhukov \\& Auch\\`ere (2004) and by Delanee (2000). However, this relationship and the nature of EIT waves remains elusive. A CME event on 5 March 2000 has been observed simultaneously by different instruments on board SOHO (EIT, LASCO and UVCS) with different signatures of wave features. This event provides a unique opportunity to establish a relationship between different phenomena such as the CME driven shock wave, streamer deflection and the EIT waves. Observations and data analysis are described in \\S2 and results and conclusions are presented in \\S3. ", "conclusions": "We studied the relationship between different coronal waves (EIT wave, streamer deflection and a CME-driven shock wave) generated by the CME event on 5 March 2000 as observed by different instruments on board SOHO (EIT, LASCO and UVCS). LASCO-C2 images show a deflection propagating outward at $\\sim260$~km~s$^{-1}$ (projected on the plane of the sky) in a remote streamer located approximately at $10^\\circ-15^\\circ$ CCW from north pole. The streamer kink was first seen in LASCO-C2 FOV ($\\sim2.5$~R$_{\\sun}$) at about 17:30~UT. No evidence for any CME material which could be the origin of such a deflection. The speed of the leading edge of the CME was $\\sim860$~km~s$^{-1}$, which is sufficient to generate a shock wave. Therefore, this propagating kink in the streamer provides strong evidence for a CME-driven shock wave in the corona as suggested by Sheeley et al. (2000). UVCS spectra show excessive broadening in \\ion{O}{vi} lines and intensity enhancement in the hot lines of \\ion{Si}{xii} 520~{\\AA} and \\ion{Mg}{x} 625~{\\AA}. These are clear evidence for a CME-driven shock wave (see RMRC00-05). The analysis of the intensity modulation along the slit also reveals the propagation direction of the wave, North-East, that is the same as the kinking streamer. UVCS observations show that the shock wave would have reached the UVCS slit around 16:45~UT. EIT difference images show evidence for an EIT wave front propagating North-East with speed of $\\sim55$~km~s$^{-1}$ (the solar disk shape not taken into account) up to the north polar hole where the propagation has stopped, satisfying the properties of EIT waves. The propagation of the wave in the South-East direction seems to be disabled by the presence of the complex active region. The EIT wave first appeared at 16:10~UT and disappeared gradually after 16:58~UT. On one hand, temporal and spatial correlations between the events observed by UVCS and LASCO suggest that they are basically different manifestations of the same phenomenon. Measurements of the speed in different part of CME envelopes reveal that noses of CMEs travel faster than their flanks (Sheeley et al. 2000). Therefore we anticipate that the propagating shock would have a similar property and would have speed higher in front of the CMEs and lower in the flanks. Different speed in the different parts would provide evidence for the anisotropic propagation of associated shock waves. On the other hand, the EIT wave observed in the low corona was significantly slower the UVCS and LASCO events. A possible explanation for this is the opening of the magnetic flux system due to the expulsion of the CME rather than the manifestations of the CME driven shock wave as suggested by Chen et al. (2005a,b) based on MHD modeling and by Delanee (2000) and by Zhukov \\& Auch\\`ere (2004). Moreton waves, which are observed in the chromosphere are interpreted as counterpart of CME driven shock waves (e.g., Chen et al. 2005a). Unfortunately, we did not have H${\\alpha}$ high cadence observations for this event, which prevented a comparison with a chromospheric Moreton wave. However, the semi-circular wavefront was moving in the same direction as the coronal shock wave. This supports the hypothesis that all three features are linked to each other. They are generated by the same source, propagate in the same direction and have good time and spatial correlations. The only apparent problem resides in the propagation speed. However, note that the three wave-like structures do not share similar physical conditions. One of these is the space and its characteristics. The speed of Alfv\\'en waves is given by $\\displaystyle{V_A=\\frac{B}{\\sqrt{4\\pi\\rho}}}$, where $B$ is the magnetic field strength and $\\rho$ is the density. Depending on models, the density drops by 3 to 4 orders of magnitude or more between the very sparse corona and 2.0-3.0~R$_{\\sun}$, where the magnetic field drops by an order of magnitude or more (assuming the the coronal field is nearly potential and thus drops as $\\displaystyle{r^{-2}}$; see Altschuler \\& Newkirk 1969). While the high corona is increasingly less dense with increasing altitude the corresponding Alfv\\'en speed tends to increase. This is not the case in the lower corona where the plasma density is higher and thus is characterized by a smaller Alfv\\'en speed. This may explain at least partially the difference in speed. Another parameter is the propagation direction of the waves. The CME-driven shock wave is mainly propagating parallel to the magnetic field lines. However, the wave propagating on the solar disk does not share this property and propagates across. It encounters different magnetic structures that may allow for energy leakage which might consequently damp and slow down the wave (see for instance De Pontieu et al. 2004 on energy leakage of the acoustic oscillation $p$ modes through flux tubes with high inclinations to the chromosphere which contributes to the heating of this layer). A good illustration for that is given by active regions through which EIT waves could not propagate and also across polar holes. We think that energy carried by the wave is progressively leaked to different magnetic structures which may heat and accelerate the plasma in these structures. This interpretation needs to be carried further and deeper through additional analysis of other observational examples and also by numerical simulations." }, "0801/0801.0599_arXiv.txt": { "abstract": "% In this contribution, I present a simplified overview of the evolution of the disk galaxy population since $z=1$, and a brief discussion of a few open questions. Galaxy evolution surveys have found that the disk galaxy population forms stars intensely at intermediate redshift. In particular, they dominate the cosmic star formation rate at $z<1$ --- the factor of ten drop in cosmic average comoving star formation rate in the last 8 Gyr is driven primarily by disk physics, not by a decreasing major merger rate. Despite this intense star formation, there has been little change in the stellar mass density in disk galaxies since $z=1$; large numbers of disk galaxies are being transformed into non-star-forming spheroid-dominated galaxies by galaxy interactions, AGN feedback, environmental effects, and other physical processes. Finally, despite this intense activity, the scaling relations of disk galaxies appear to evolve little. In particular, as individual galaxies grow in mass through the formation of stars, they appear to grow in radius (on average, the population grows inside-out), and they appear to evolve towards somewhat higher rotation velocity (i.e., mass is added at both small and large radii during this inside-out growth). ", "introduction": "The properties and evolution of the disk galaxy population are a critical diagnostic of the galaxy formation process in a dark matter-dominated Universe. Galaxy disks, formed naturally as a result of a collapse in which at least some of the angular momentum is conserved \\citep{fall80}, are relatively fragile beasts. Interactions with low-mass halos are likely to thicken the stellar disk \\citep[e.g.,][]{toth92,benson04} and/or lead to warps and lopsidedness; interactions with galaxies/halos with masses within a factor of a few of the galaxy mass look likely to disrupt the stellar disk entirely \\citep[e.g.,][]{koda08}. Thus, the evolution and properties of the disk galaxy population are not only a fascinating puzzle in their own right, but also give insight into those environments which are the least affected by the characteristic pummeling that galaxies in $\\Lambda$CDM seem to receive. In the seven years since the first Vatican disk galaxies meeting, a number of the basic observational features of the evolution of the disk galaxy population have come into place. In this contribution, I present a brief overview of what I consider to be some of the most important features of the evolution of the disk galaxy population since $z=1$ (roughly the last 7-8 billion years), and pose a few questions that I had after this conference. These features fall under three broad themes: star formation in disks, stellar mass in disks, and disk galaxy scaling relations. In what follows, in the spirit of a review, I will deliberately over-simplify the observational results somewhat in order to clarify the three basic messages of this contribution. ", "conclusions": "In this contribution, I have presented an over-simplified view of the evolution of the disk galaxy population: \\begin{itemize} \\item Disk galaxies form stars intensely since $z=1$. They dominate the cosmic average SFR since $z=1$; the factor of 10 decline in cosmic SFR since $z=1$ is primarily driven by disk galaxy physics (gas consumption, decreased accretion from the cosmic web), not changes in the galaxy merging rate. \\item Despite this intense star formation, the disk galaxy population does not grow in mass since $z=1$. Large numbers of star-forming disks are being transformed into non-star-forming spheroid-dominated galaxies. \\item As disk galaxies gain mass through star formation, they grow in radius (i.e., the population grows inside-out) and rotation velocity (i.e., mass is added at small and large radii during this inside-out growth). \\end{itemize} A final word. This picture of a dynamic disk galaxy population has important implications for how we interpret the results of our observations. Statements such as 'massive disk galaxies are in place at $z=1$ and appear to evolve little to the present day' are incorrect. Some of these massive disk galaxies have become bulge-dominated at the present day (and are therefore not in the $z=0$ control disk galaxy sample), some are in (rare) even more massive disk galaxies. This has implications for how one should interpret a variety of observations: bar fractions, metallicity--mass relations, mass--radius and mass--rotation velocity relations. While some studies specifically address such issues (e.g., \\citealp{kassin06}), it is important to bear the (observed!) dramatic evolution of the disk galaxy population in mind when interpreting scaling relations and properties of the evolving disk galaxy population." }, "0801/0801.3848_arXiv.txt": { "abstract": "I review recent progresses in the dynamics and the evolution of self-gravitating accretion discs. Accretion discs are a fundamental component of several astrophysical systems on very diverse scales, and can be found around supermassive black holes in Active Galactic Nuclei (AGN), and also in our Galaxy around stellar mass compact objects and around young stars. Notwithstanding the specific differences arising from such diversity in physical extent, all these systems share a common feature where a central object is fed from the accretion disc, due to the effect of turbulence and disc instabilities, which are able to remove the angular momentum from the gas and allow its accretion. In recent years, it has become increasingly apparent that the gravitational field produced by the disc itself (the disc's self-gravity) is an important ingredient in the models, especially in the context of protostellar discs and of AGN discs. Indeed, it appears that in many cases (and especially in the colder outer parts of the disc) the development of gravitational instabilities can be one of the main agents in the redistribution of angular momentum. In some cases, the instability can be strong enough to lead to the formation of gravitationally bound clumps within the disc, and thus to determine the disc fragmentation. As a result, progress in our understanding of the dynamics of self-gravitating discs is essential to understand the processes that lead to the feeding of both young stars and of supermassive black holes in AGN. At the same time, understanding the fragmentation conditions is important to determine under which conditions AGN discs would fragment and form stars and whether protostellar discs might form giant gaseous planets through disc fragmentation. ", "introduction": "Disc-like or flattened geometries are very common in astrophysics, from the large scale of spiral galaxies down to the small scales of Saturn's rings. In both these examples, the discs are characterized by a prominent structure either in the form of a spiral structure (in the case of galaxy discs) or in the form of gaps and spirals on small scales (in the case of Saturn's rings). The dynamics underlying the development of such structures is determined by the propagation of density waves, and the important role of the disc's self-gravity in their development has been clearly recognized (see, for example, \\cite{bertinbook}). In the above examples the system is either{\\it collisionless}, such as in the case of spiral galaxies, where the dominant component is constituted of stars, or {\\it particulate}, such as in the case of Saturn's rings, where the dominant components (mainly rocks and pebbles) do undergo inelastic collisions, but the system cannot be simply described in terms of hydrodynamics. In the last thirty years increasing attention has been given to {\\it fluid} discs, where the dynamically active component is gaseous. Here, dissipative effects, associated with friction or viscosity, can significantly affect the dynamics of the disc. Through dissipation the fluid elements in the disc can lose their energy or, more fundamentally, their angular momentum, as I describe in more detail below, and can fall towards the bottom of the potential well, hence accreting on to a central gravitating body. Such a system, where a disc feeds a central object through accretion, under the effect of viscous forces, is called an {\\bf accretion disc}. Accretion discs are found in very different contexts and over a wide range of physical scales. On the largest scale, they are one of the main physical ingredient to power the central engine of Active Galactic Nuclei (AGN), through accretion on to a supermassive black hole. The mass of the central black hole ranges from $10^6M_{\\odot}$ to $10^9M_{\\odot}$ and the accretion discs can extend out to large distances, of the order of about 1 pc, where rotating gaseous discs have often been detected through water maser emission \\cite{miyoshi95,greenhill97,kondratko06}. On the galactic scale, accretion discs around compact objects, such as white dwarfs, neutron stars and stellar mass black holes, are often found in binary systems, where the compact object (with a mass of the order of $1M_{\\odot}$) is fed by material outflowing from the companion. Historically, this has been one of the first contexts where accretion disc theory has found widespread application (a detailed review with an emphasis on accretion in galactic binary system and AGN can be found in the textbook by Frank, King and Raine \\cite{franck}). Another class of objects where accretion discs play an important role are Young Stellar Objects (YSO). Here, the central accreting object is a young protostar, which receives most of its mass from a surrounding protostellar discs. In this case, protostellar discs play also another key role, as the site where planet formation takes place. Understanding the dynamics of the disc and the development of the various instabilities that might occur in it, and possibly lead to turbulence, is therefore essential if one wants to understand some properties of our own Solar system, as well as those of extra-solar planets. A thorough description of the current observational state of protostellar discs can be found, for example, in the recent Protostars and Planets V book \\cite{PPV}. As for the case of spiral galaxies, also for accretion discs the effects of the disc self-gravity can be very important. Indeed, also in this case the development of gravitational instabilities can lead to the formation of grand design spiral structures, which deeply affect the structure and the dynamics of the disc. The analysis of such instabilities naturally reveals several similarities between the case of spiral galaxies and that of accretion discs, but also some important differences. First of all, as mentioned above, accretion discs are fluid and, as such, they are intrinsically dissipative. Indeed, the evolution of gravitationally unstable accretion discs, as I will discuss below, is strongly dependent on the gas cooling rate. On the other hand, gravitationally unstable discs of stars, in the absence of any gas, have a natural tendency to dynamically heat up. Thermal energy, or disordered kinetic energy, plays a key role in stabilizing self-gravitating discs. However, while for stellar discs this energy is mostly in the form of disordered stellar motions, with the possibility of an anisotropic velocity dispersion, for gaseous discs the disordered kinetic energy content is mostly in the form of thermal energy. Furthermore, for gaseous discs it is expected that the role of resonances is much reduced with respect to a collisionless system. Finally, while for galaxy discs the disc mass can be a substantial fraction of the total mass of the system and therefore give a large contribution to the gravitational potential, in the case of accretion discs the disc mass is often much smaller than the mass of the central object (with some important exceptions, described below). Historically, accretion disc theory has concentrated on the non self-gravitating case, the effects of self-gravity having only been discussed occasionally \\cite{pacinski78,kolikalov79,linpringle87,linpringle90}. In the last ten years, on the other hand, the important role of the disc self-gravity has been clearly recognized, partly due to improved observations, which have shown that in several observed systems (on all scales, from AGN to protostars) the disc mass can be high enough to have a dynamical role, and partly due to the increased computational resources, which have allowed a detailed numerical investigation of the development of gravitational instabilities in the non-linear regime. However, the early papers mentioned above already reveal the main directions around which research in this field would have later developed. Currently, most of the interest revolves around the three issues of self-regulation \\cite{pacinski78}, disc fragmentation \\cite{kolikalov79} and angular momentum transport induced by gravitational instabilities \\cite{linpringle87,linpringle90}. This again reveals a difference with respect to the galaxy discs case, where most of the attention has been traditionally given to the morphology induced by self-gravity. In this paper, I present an overview of the dynamics of self-gravitating accretion discs. Accretion disc theory is a very vast topic and has been treated extensively elsewhere \\cite{franck,pringle81}, so I will not repeat it here. I will rather summarize the most salient general features, with particular emphasis on those features which are most affected by the disc self-gravity. Similarly, a detailed description of all the different astrophysical systems where self-gravitating accretion discs play a role (protostellar discs, AGN,...) is obviously beyond the scope of this review. The structure of the paper is as follows. In section \\ref{sec:dynamics} I describe the basic equations that determine the evolution of accretion discs and the main effects of the disc self-gravity. In section \\ref{sec:GI} I discuss the development of gravitational instabilities, and present some models which incorporate in a simple way the salient physics behind their development. In section \\ref{sec:simulations} I review the current state of numerical simulations of gravitationally unstable gaseous discs. In section \\ref{sec:transport} I focus on the transport properties induced by gravitational instabilities, while in section \\ref{sec:fragmentation} I describe the related issue of disc fragmentation. Finally, in the last two sections I will present some examples of observed systems where the concepts described in this paper find a natural application, and in particular, in section \\ref{sec:examples} I will discuss the impact of self-gravity in the dynamics of protostellar discs with some application related to planet formation. In section \\ref{sec:AGN} I focus on the impact of self-gravity in AGN discs also showing how self-gravitating accretion discs might have played an important role at high red-shifts, by allowing the formation of the seeds of supermassive black holes. ", "conclusions": "\\label{sec:conclusions} As we have seen, self-gravitating accretion discs are an important aspect of the modeling of several systems, and thus play a key role in modern astrophysics. The influence of the disc self-gravity extends from the very small scales, associated with the formation of planetary systems, up to the large scales, associated with the formation and the feeding of supermassive black holes in the nuclei of galaxies. While clearly the specific properties of systems associated with such hugely different scales are going to differ substantially, there is yet a unifying framework related to the scale-free nature of the gravitational force. Because of self-gravity, accretion discs can develop gravitational instabilities. These are going to deeply affect the structure and the evolution of the accretion disc. Firstly, we have seen that the instability usually takes the form of a regular spiral structure, that can efficiently redistribute the disc angular momentum and thus allow the accretion process to take place. In this respect the role of the disc self-gravity is complementary to that of other instabilities, such as magneto-hydrodynamic (MHD) instabilities \\cite{balbusreview,balbus03}. Indeed, while MHD instabilities are likely to be the most important source of angular momentum transport in the inner and hotter parts of the disc, it is not clear whether they are able to provide the necessary torques in colder environments, where the ionization level of the gas is low and the coupling to the magnetic field is weaker. This is, for example, the case of the outer parts of protostellar discs \\cite{fromang02}. On the other hand, such cold parts of the discs are naturally subject to gravitational instabilities. It is then possible to envisage a situation where gravitational instabilities are responsible for bringing the accreting material from large distances (which might be of the order of tens of AU for protostellar discs, and a fraction of a parsec for AGN discs) into the inner disc, where MHD instabilities take over and are responsible for the ultimate deposition of the matter onto the central object. In this respect, it would be important to develop models which include both the effects of MHD induced turbulence and that of gravitational instabilities. Semi-analytical models along these lines have been discussed \\cite{kratter07}, but it is still unclear what would be the interplay between magnetic and gravitational instabilities in the general case \\cite{fromang04a,fromang04b}. A second aspect related to gravitational instabilities is the natural tendency associated with self-gravity to collapse and produce gravitationally bound objects. Disc fragmentation is a possible outcome of the onset of gravitational instabilities. As discussed above, it is now well established that the conditions leading to fragmentation are essentially related to the disc thermodynamics and to the cooling rate. From the modeling point of view, fragmentation has a twofold aspect. On the one hand, if one is interested in producing smaller objects within the disc, it offers a natural way to accomplish the task. On the other hand, there is the danger that most of the disc mass would end up in the fragments, with very little being accreted onto the central object. In this respect, fragmentation is a dangerous outcome, which might inhibit accretion. We have seen that the biggest difference between protostellar discs and AGN discs with respect to self-gravity lies indeed in their thermal properties, and in particular in the fact that AGN discs are expected to be cold (in the sense that $H/R\\ll 1$) and with a short cooling timescale (in the sense that $\\tau_{\\rm cool}\\Omega\\ll 1$), while protostellar discs are hot ($H/R\\approx 0.1$) and with a long cooling timescale ($\\tau_{\\rm cool}\\Omega\\gg 1$, at least within 100 AU). The behaviour of these two kind of systems in this respect is then significantly different. Most protostellar discs are not expected to fragment, except perhaps in their outermost regions. This is the main problem affecting models of planet formation which rely on direct disc fragmentation. At the other end of the scales that we have considered, we see that AGN discs instead are likely to fragment. This is good news in view of forming young massive stars in the vicinity of the central supermassive black hole, as are observed, for example, close to the black hole in our Galactic Center. On the other hand, we have discussed above the problems that might arise from disc fragmentation in this context, related to the feeding of the black hole. The last 10-15 years have witnessed a significant improvement of our understanding of the evolution of self-gravitating accretion discs. This has been made possible through the interplay between simple analytical models and the results of detailed numerical simulations of the disc hydrodynamics. The big advances in computing power in the last decade have made it possible to run such numerical simulations with unprecedented resolution, and progressively including the effects of several physical processes in more detail. In this paper, emphasis has been put on the important role of the gas cooling rate in determining the outcome of the instability. This has been often introduced in a simplified way in the simulations, which has led to a valuable investigation of the different regimes, in a somewhat academic way. The theoretical challenge in this respect is to introduce in the numerical codes a more realistic cooling function, in order to understand the behaviour of actual systems. For example, most of the current debate on the applicability of the fragmentation scenario for planet formation is ultimately due to our ignorance of the actual thermodynamics at work in protostellar discs. The difficulty here is to treat numerically the radiative transfer within very optically thick media. Only very recently have numerical codes implemented radiation physics and, although still preliminary, some results are starting to emerge \\cite{boley06,mayer07}. Codes which couple radiative transfer with the disc hydrodynamics are also desired to include the possibly important effect of irradiation from the central object, which has mostly been neglected in the simulations run so far (with some exceptions \\cite{cai06}). If we look at the problem of star formation within AGN accretion discs, here the issue is ultimately whether accretion can proceed at significant rates even if the disc fragments. In other words, it is still unclear how much gas needs to be turned into stars before fragmentation is quenched. Here, it would be important to include in the models the energy feedback arising from the forming stars. Even in this context then, the key ingredient to be added into numerical codes is a proper treatment of radiative transfer, coupled with hydrodynamics." }, "0801/0801.4373_arXiv.txt": { "abstract": "We discuss recent models on the evolution of massive stars at very low metallicity including the effects of rotation, magnetic fields and binarity. Very metal poor stars lose very little mass and angular momentum during the main sequence evolution, and rotation plays a dominant role in their evolution. In rapidly rotating massive stars, the rotationally induced mixing time scale can be even shorter than the nuclear time scale throughout the main sequence. The consequent quasi-chemically homogeneous evolution greatly differs from the standard massive star evolution that leads to formation of red giants with strong chemical stratification. Interesting outcomes of such a new mode of evolution include the formation of rapidly rotating massive Wolf-Rayet stars that emit large amounts of ionizing photons, the formation of a long gamma-ray bursts and a hypernovae, and the production of large amounts of primary nitrogen. We show that binary interactions can further enhance the effects of rotation, as mass accretion in a close binary spins up the secondary. ", "introduction": "Massive stars affect the evolution of the early universe in a number of ways. They are believed to be the main source of reionizing photons at high redshift, and their explosions provide large amounts of energy into the surrounding medium. Most elements heavier than helium begin to exist in the early universe due to nucleosynthesis in massive stars. The history of star formation and the evolution of galaxies in the early universe are therefore closely related to such feedback from massive stars. This has motivated many theoretical studies on the evolution of massive stars of the first and second generations, which are characterized by zero/low metallicity. Mass loss due to stellar winds from massive stars is mainly driven by metal lines (Castor, Abbott \\& Klein~1975; Pauldrach, Puls \\& Kudritzki R.P.~1986; Vink, de Koter \\& Lamers~2001). As a result, metal poor massive stars are expected to lose only a small amount of mass and angular momentum and to be kept rotating rapidly, during their life times (see Ekstr\\\"om and Meynet, however, in this volume). Rotation -- one of the primary factors to determine the evolution of massive stars -- may thus play an even more important role in the evolution of massive stars in the early universe, compared to the case of galactic metal-rich massive stars. For instance, the high ratio of nitrogen to carbon abundance observed in extremely metal poor stars may be related to chemical mixing due to rotationally induced instabilities in massive stars in the early universe (Chiappini et al.~2006). Stellar cores could also more easily retain a large amount of angular momentum at lower metallicity, resulting in the abundant production of energetic supernovae and gamma-ray bursts (Yoon \\& Langer~2005; Woosley \\& Heger~2006; Yoon, Langer \\& Norman~2006). Detailed numerical simulations of the evolution of massive stars including the effects of rotation involve many uncertain physical processes such as the transport of angular momentum and chemical species due to rotationally induced instabilities. Studies by Langer et al. (1999), Heger, Langer \\& Woosley (2000) and Hirschi, Meynet \\& Maeder (2004) indicate that their adopted angular momentum transport mechanisms (Eddington Sweet circulations, shear instability and baroclinic instability) are too inefficient to explain the observed spin rates of white dwarfs and young neutron stars: their models predict one or two orders of magnitude higher spin rates than the observed values. More recent models adopting the prescription by Spruit (2000) for the Tayler-Spruit dynamo in differentially rotating radiative layers are shown to be more consistent with observations in terms of the spin rates of stellar remnants (Heger, Woosley \\& Spruit~2005; Suijs et al. in prep.). The validity of the Talyer-Spruit dynamo has recently been challenged by several authors (Denissenkov, Pavel, \\& Pinsonneault~2007; Zahn, Brun \\& Mathis, S.~2007) but it seems clear that an efficient braking mechanism that is comparable to what Spruit suggests is needed to explain observations. Here we present recent models of both single and binary massive stars at very low metallicity including the Tayler-Spruit dynamo as well as other effects of rotation such as rotationally induced chemical mixing, enhanced mass loss near the break-up velocity and tidal interactions. We suggest that rotation could lead to very different types of massive star evolution at very low metallicity, compared to the case of metal rich stars. Implications for supernovae and gamma-ray bursts, and massive star feedback in the early universe are briefly discussed. ", "conclusions": "We still do not fully understand the details about the effects of rotation such as transport processes of angular momentum and chemical species due to magnetic torques, Eddington Sweet circulations and other possible rotationally induced hydrodynamic instabilities in stars, which are important ingredients in recent stellar evolution models. Improved treatments of the rotation-related physics are thus needed for future studies, in connection with observational tests and multi-dimensional simulations (e.g. Brott et al. in this volume; Talon~2007). However, above discussions clearly indicate that the evolution of metal poor massive stars much depends on rotation. Addressing the role of the massive star feed back in the early universe including the effects of rotation and binarity will be an exciting but challenging subject for the next decade. In particular, future observational studies on massive star populations (e.g. WR/O ratio) in metal poor galaxies and supernovae/GRBs at high redshift may provide strong constraints for theoretical models, as discussed in Yoon, Langer \\& Norman~(2006) in greater detail." }, "0801/0801.3286_arXiv.txt": { "abstract": "The observed properties of galaxies vary with inclination; for most applications we would rather have properties that are independent of inclination, \\emph{intrinsic} properties. One way to determine inclination corrections is to consider a large sample of galaxies, study how the \\emph{observed} properties of these galaxies depend on inclination and then remove this dependence to recover the \\emph{intrinsic} properties. We perform such an analysis for galaxies selected from the Sloan Digital Sky Survey which have been matched to galaxies from the Two-Micron All Sky Survey. We determine inclination corrections for these galaxies as a function of galaxy luminosity and Sersic index. In the $g$-band these corrections reach as as high as 1.2 mag and have a median value of $0.3$ mag for all galaxies in our sample. We find that the corrections show little dependence on galaxy luminosity, except in the $u$ band, but are strongly dependent on galaxy Sersic index. We find that the ratio of red-to-blue galaxies changes from 1:1 to 1:2 when going from observed to intrinsic colors for galaxies in the range $-22.75 < M_K < -17.75$. We also discuss how survey completeness and photometric redshifts should be determined when taking into account that observed and intrinsic properties differ. Finally, we examine whether previous determinations of stellar mass give an intrinsic quantity or one that depends on galaxy inclination. ", "introduction": "In our search to understand the formation and evolution of galaxies, some of our primary tools are measurements of the distribution of galaxy properties and relationships among these properties. From observations of these quantities at different redshifts we can deduce the nature of galactic evolution and by comparing them to the properties of dark matter halos we can constrain models of galaxy formation. The measurement of galactic distributions includes, but is not limited to: the galaxy luminosity function \\citep{hubb:36}, the galaxy correlation function \\citep[the distribution of galaxies' spatial separations]{ph:74}, the galaxy velocity function \\citep{gonz:00}, and the distributions of galaxy sizes \\citep{chol:85}, surface brightnesses \\citep{free:70}, colors \\citep{baum:59,faber:73}, metalicities \\citep{oste:70} and star formation rates \\citep{td:80}. Relationships between galaxy properties include, the Tully-Fisher relation \\citep{tf:77}, the Faber-Jackson \\citep{fj:76} and fundamental plane \\citep{dd:87,dres:87} relations, the luminosity-size relation \\citep{korm:77}, the luminosity-metalicity relation \\citep{faber:73, lequ:79} and the density-morphology relation \\citep{dres:80}. However, with the notable exception of the Tully-Fisher relation these distributions and relations are traditionally measured in terms of the {\\it observed} properties of galaxies. That is, the measurements used are K-corrected and corrected for foreground dust extinction, but no correction is attempted to compensate for the viewing angle from which the galaxies are observed. In contrast, the Tully-Fisher relation is not a relationship between a galaxy's \\emph{observed} luminosity and rotation velocity, but a relation between a galaxy luminosity and rotation velocity \\emph{corrected for inclination}. The inclination correction attempts to recover the intrinsic properties of a galaxy and not properties that are measured because of the particular angle from which the galaxy is viewed. Spiral galaxies are observed to have redder colors when their disks are more inclined, which is expected if the inclination increases the amount of dust that light traverses when emitted from the galaxy. Clearly, we would prefer to measure all galactic distributions and relationships in terms of intrinsic galaxy properties instead of observed ones. The comparison of theory to observations is complicated and often done incorrectly because of confusion between observed and intrinsic galaxy properties. Early semi-analytic models were unable to match both the galaxy luminosity function and the Tully-Fisher relation in part because they failed to take into account that the first is observed luminosity while the second is intrinsic luminosity \\citep{sp:99}. Also, when comparing galaxies at different redshifts we would like to be able to distinguish between evolution in their stellar populations and changes in their dust properties. Furthermore, inclination effects are of great help in understanding the nature of dust in galaxies. Theoretical modeling of attenuation in galaxies is complicated because it not only depends on the properties of dust, which seem to vary between galaxies, but also on how the dust is distributed and mixed with stars. By determining the intrinsic properties of galaxies we also learn how those properties change as a function of galaxy inclination and therefore some properties of the dust distribution. To determine intrinsic properties we need to know how a galaxy's properties change as a function of its inclination. This is different then removing the effects of dust and dust is still present for a face-on galaxy. There are a number of approaches for addressing this issue each of which has its own merits and disadvantages. One approach is to solve for an inclination correction that minimizes the scatter in the Tully-Fisher relation \\citep[e.g.,][]{verh:01}, which assumes that the scatter in this relation should be as small as possible. Another method is to fit stellar population models to the SED of a galaxy and then assume that any discrepancies are caused by dust \\citep[e.g.,][]{kauf:03}. A third is to observe background objects behind a foreground galaxy to get a direct measure of the extinction through the galaxy \\citep[e.g.,][]{berl:97,holw:05}, but this is difficult to do for more than a handful of cases. Finally, one can simulate the radiative transfer through a galaxy \\citep[e.g.,][]{rocha:07}, assuming one knows the distribution and scattering properties of the dust. The approach we explore here is somewhat simpler in that it assumes no knowledge of stellar population or dust properties. Instead, the main assumption is that a galaxy's properties should be independent of inclination. Thus any statistical correlation between a galaxy property and inclination can be attributed to dust and the inclination correction is whatever makes the observed correlation go away. This procedure has been applied a number of times \\citep{giov:94,giov:95,tully:98,mgh:03,shao:07}. In this paper we greatly expand upon this method by applying it to 10,340 galaxies taken from the Sloan Digital Sky Survey \\citep[SDSS,][]{york:00} with accompanying infrared magnitudes from the Two-Micron All Sky Survey \\citep[2MASS,][]{skru:06}. It is important to have galaxies with near infrared photometry because the effects of attenuation are minimized in these wavebands \\citep{bd:01}. In a subsequent paper we will extend the analysis performed here to the full SDSS galaxy catalog. We describe the method for determining inclination corrections in \\S\\ref{sec:method}. In \\S\\ref{sec:data} we describe the sample we will use and discuss some of the properties of galaxies in this sample. In \\S\\ref{sec:results} we determine inclination corrections using our sample and compare our results to other determinations in \\S\\ref{sec:compare}. In \\S\\ref{sec:changes} we discuss how consideration of intrinsic properties can change our conclusions about the distribution of galaxy properties focusing on the color-magnitude diagram. We also comment on the effect on survey completeness, stellar masses and photometric redshifts. \\S\\ref{sec:conc} contains our conclusions and some discussion of future directions. ", "conclusions": "\\label{sec:conc} This paper has focused on the intrinsic properties of galaxies, which are routinely determined for Tully-Fisher studies \\citep[e.g.,][]{piza:07}, but not for other statistical studies of galaxies. Intrinsic properties are invariant under changes in viewing angle, unlike observed properties which would change if we could view a galaxy from a different vantage point. Our main goal in this paper has been to clarify the difference between observed and intrinsic properties and to suggest how observed properties can be converted into intrinsic ones. The method we use in this paper is to identify an observed galaxy property that shows a correlation with axis ratio and then apply the necessary correction to remove this correlation. We find that both color and size show correlations with axis ratio. We are able to remove these correlations using simple linear formula that depend on $K$-band magnitude and Sersic index. We therefore can construct a galaxy catalog with intrinsic value for galaxy size and magnitude. There are many distributions and relationships that should be reconsidered in terms of the intrinsic properties instead of the observed ones. We can not cover all of these in a paper of this length but we highlight a few points to suggest how things may change. We focus on the color magnitude diagram as an example. The observed color magnitude diagram shows a number of galaxies redder than the red sequence and a lack of bright blue galaxies. When we plot the intrinsic color magnitude diagram these ultra red galaxies are mostly removed and there are many more bright blue galaxies. We find that the ratio of blue to red galaxies changes from 1:1 to 2:1 for galaxies with absolute luminosities $-23.75 \\ge M_K > -17.75$, a significant change. In a following paper, we apply the insights we have gained here to producing inclination corrections for the full SDSS catalog. Having intrinsic properties for this flux limited catalog will then allow us to address how volume densities are effected when going from observed to intrinsic quantities." }, "0801/0801.4529_arXiv.txt": { "abstract": "We perform a study of cosmic evolution with an equation of state parameter $\\omega(t)=\\omega_0+\\omega_1(t\\dot H/H)$ by selecting a phenomenological $\\Lambda$ model of the form, $\\dot\\Lambda\\sim H^3$. This simple proposition explains both linearly expanding and inflationary Universes with a single set of equations. We notice that the inflation leads to a scaling in the equation of state parameter, $\\omega(t)$, and hence in equation of state. In this approach, one of its two parameters have been pin pointed and the other have been delineated. It has been possible to show a connection between dark energy and Higgs-Boson. ", "introduction": "Cosmological research is mainly concerned with time (and in some cases space as well) evolution of various physical parameters like scale factor, Hubble parameter, matter-energy density etc. Along with these parameters, in recent years a new physical entity $\\Lambda$ has resurrected in the foreground of cosmology. In fact, $\\Lambda$ has become an essential part of the field equations of Einstein after some observational results \\citep{Riess1998,Perlmutter1999} indicated towards an accelerating Universe. It is believed by most of the physicists that the cosmological parameter $\\Lambda$ is responsible for driving the present acceleration because it can exert negative pressure. Moreover, due to some fine-tuning problem (known as cosmological constant problem), $\\Lambda$ is regarded as a variable quantity rather than a constant. Now, in order to specify exact time-dependence of the unknown physical quantities including $\\Lambda$, one has to take recourse of a relationship between cosmic pressure $p$ and matter-energy density $\\rho$ involving the equation of state parameter $\\omega$. Mathematically speaking, one variable quantity can depend on the product of two other variable quantities. So, one may construct $\\omega$ as a function of time, red-shift or scale factor \\citep{Chevron2000,Zhuravlev2001,Peebles2003}. In fact, values of $\\omega$ at different stages of cosmic evolution suggest that it may evolve with time. As an instance, for the present pressure-less Universe, the value of $\\omega$ is considered as zero, whereas its value was $1/3$ in the early radiation dominated Universe. However, it is convenient to consider $\\omega$ as a constant quantity because observational data can hardly distinguish between a varying and a constant equation of state \\citep{Kujat2002,Bartelmann2005}. Here some useful limits on $\\omega$ as appeared from SNIa data are $-1.67 <\\omega < -0.62$ \\citep{Knop2003} whereas refined values were indicated by the combined SNIa data (with CMB anisotropy) and galaxy clustering statistics which is $-1.33 < \\omega < -0.79$ \\citep{Tegmark2004}. As stated above, $\\omega$ may have a functional relationship with scale factor or cosmological redshift. In connection to redshift it may depend linearly, $\\omega(z) = \\omega_o + {\\omega}^{\\prime} z$, where ${\\omega}^{\\prime}= (d\\omega/dz)_{z=0}$ \\citep{Huterer2001,Weller2002} or it may have a non-linear relationship as $\\omega(z) = \\omega_o + {\\omega}_1 z/(1+z)$ \\citep{Polarski2001,Linder2003}. This suggests for a simple form \\begin{eqnarray} \\label{omegaq} \\omega(t)= \\omega_0 + \\omega_1 (t \\dot H/H), \\end{eqnarray} which has got an explicit time dependence that disappears with the condition, $t\\dot H =H$. Using above proposition, we explore the physical features of different stages of cosmic evolution, viz., linearly expanding and inflationary Universes. For this, a phenomenological $\\Lambda$ model is selected to solve the Einstein field equations. There are mathematically motivated works~\\citep{Ray2007,Mukhopadhyay2005,Mukhopadhyay2007a,Mukhopadhyay2007b}, wherein several phenomenological $\\Lambda$ models have been investigated for time-dependent $\\omega$. ", "conclusions": "We have discussed two Universes: (i) a linearly expanding Universe from its very beginning, (ii) and also the Universe like ours, which has gone through an inflation at its very early stage followed by a linear expansion later. We notice that these two kind of Universes, which are direct consequence of our proposition (equation~\\ref{omegaq}), are represented by the same set of equations with a translational shift in the equation of state parameter in the latter case compared to the former. In both the cases, $a(t)=1$ demands $E=1$, which applies a constraint on the equation of state parameter. For the inflationary Universe, we have pin pointed $\\omega_0=-1/3$ and have delineated the other parameter with a range $-2/3 <\\omega_1< -0.46$. We observe that former is a special case of the latter with $\\omega_0=-1/3$ and $\\omega_1$=0. Any other value of $\\omega_0$ would invoke a non-linear behaviour in $a(t)$ through $E$. The effect of the variation of $\\omega_0$ on $p$ is presented in Figure~\\ref{fig2} for a constant $\\omega=\\omega_0+\\omega_1=-0.80$ obtained by adjusting $\\omega_1$ accordingly. The $\\omega_1$ has nothing to do with $E$ and hence has nothing to do with $a(t)$. However, its value is a measure of translation in $\\omega$ due to inflation. The equations for $\\rho$ and $p$ involve $\\omega$ and hence would remain unchanged with its constant value. Thus, variations in curves of Figure~\\ref{fig2} is purely due to the variation in $\\omega_0$. The corresponding variations in $a(t)$ are shown in Figure~\\ref{fig3}. A negligible value of $A$ is shown to be physically possible from the viewpoint of cosmology and particle physics, which means the absence of $\\Lambda$ in the field equations. So, both from physical and mathematical point of view the nullity of $\\Lambda$ is achieved for the same $\\Lambda$ model. Again, the expression of $q$ in this case has a striking similarity with that of~\\citet{Ray2007}. This work suggests that in the late phase of the Universe, where $t\\dot H=H$, the equation of state parameter behaves as a constant. Perhaps for this reason current data cannot distinguish clearly between a time-dependent $\\omega$ and a constant one as pointed out by some workers \\citep{Kujat2002,Bartelmann2005}. Separating the entire cosmic history into two phases, it has been possible to derive the time-dependent expressions for the scale factor and the other physical parameters of each phase. It has been found that for inflationary phase, the deceleration parameter $q$ depends on time whereas for the linearly expanding phase it is constant, rather zero. This supports the opinion that $q$ has changed during the course of time~\\citep{Riess2001,Amendola2003,Padmanabhan2003}." }, "0801/0801.3543.txt": { "abstract": "{T Tauri stars exhibit variability on all timescales, whose origin is still debated. On WTTS the variability is fairly simple and attributed to long-lived, ubiquitous cool spots.} {We investigate the long term variability of WTTS, extending up to 20 years in some cases, characterize it statistically and discuss its implications for our understanding of these stars.} {We have obtained a unique, homogeneous database of photometric measurements for WTTS extending up to 20 years. It contains more than 9 000 UBV R observations of 48 WTTS. All the data were collected at Mount Maidanak Observatory (Uzbekistan) and they constitute the longest homogeneous record of accurate WTTS photometry ever assembled.} {Definitive rotation periods for 35 of the 48 stars are obtained. Phased light curves over 5 to 20 seasons are now available for analysis. Light curve shapes, amplitudes and colour variations are obtained for this sample and various behaviors exhibited, discussed and interpreted.} {Our main conclusion is that most WTTS have very stable long term variability with relatively small changes of amplitude or mean light level. The long term variability seen reflects modulation in the cold spot distributions. Photometric periods are stable over many years, and the phase of minimum light can be stable as well for several years. On the long term, spot properties do change in subtle ways, leading to secular variations in the shape and amplitudes of the light curves.} ", "introduction": "The third catalog of pre-main-sequence emission-line stars assembled by G. Herbig and collaborators includes 742 objects (Herbig \\& Bell 1988). Of these, about fifty objects were classified as weak-line T Tauri stars (WTTS).These include objects from X-ray surveys of several star-forming regions and from a survey of stars with CaII H and K emis\\-sion lines.WTTS exhibit a weak, narrow $H\\alpha$ line [W($H\\alpha$) $<$ 10\\AA], a strong LiI absorption line (6707 \\AA) with W(Li) $>$ 0.1 \\AA, and a substantial CaII H and K emission lines. The first photoelectric UBVR observations of these stars revealed periodic light variations in most WTTS (Rydgren \\& Vrba 1983; Rydgren et al. 1984; Bouvier et al. 1986, 1993, 1995; Grankin 1992, 1993, 1996; Vrba et al. 1993). Their photometric periods ranged from 1 to 10 days and were in agreement with spectroscopic estimates of their rotational velocities. It was assumed that the periodic light variations were due to the nonuniform distribution of cool photospheric spots over the stellar surface. Rotation periods are now available for some hundreds of pre-main sequence and recently arrived main sequence stars of solar-like mass in five nearby young clusters: the Orion Nebula Cluster, NGC 2264, $\\alpha$ Per, IC 2602 and the Pleiades (Mandel \\& Herbst 1991; Attridge \\& Herbst 1992; Eaton et al. 1995; Edwards et al. 1993; Choi \\& Herbst 1996; Bouvier et al. 1997a). In combination with estimates of stellar radii these data allow us to construct distributions of surface angular momentum per unit mass at three different epochs: nominally, 1, 2 and 50 Myr (see Herbst \\& Mundt 2005). Recent work is extending the age range to 200 Myr and adding many new clusters (Irwin et al. 2007). The study of long-term variations in the main param\\-eters of the light curves for young spotted stars is also of considerable interest. One would expect light-curve shapes to evolve as the spots change size, shape, temperature, or location, and a latitudinal migration of spots could cause changes in period, if the surfaces are in differential rota- tion, as in the Sun. It would, of course, be very interest- ing to find any cyclic pattern that could be interpreted as evidence for a magnetic cycle. Unfortunately, most of the photometric programs of observations of WTTS stars have been carried out episodically, in an interval from one to three months. Only one object (V410 Tau) has been in\\-vestigated in detail over several seasons (Vrba et al. 1988; Herbst 1989; Petrov et al. 1994). These authors showed that there were at least two extended and long-lived spot- ted regions on the surface of V410 Tau during several years. The evolution of the shape and amplitude of the light curve from season to season was explained by changes in the spot sizes, temperatures, and relative locations. The first systematic photoelectric BVR observations for about thirty WTTS in dark clouds in Taurus-Auriga began in 1990 at the Maidanak Observatory (Grankin et al. 1995). During 1990-1993 rotation periods were dis- covered for 12 WTTS and refined for another 9 WTTS. A fundamental result of the analysis of the photometric behavior of these WTTS was the stability of the initial epochs and rotation periods for 17 WTTS on a time scale from 2 to 4 years. In addition, there is a high detection rate of rotation periods among WTTS in the sample con- sidered. It is generally believed that the stability of the initial epochs and rotation periods over several years in- dicates that the active region in each WTTS remains on a definite meridian over this time scale. Variations in the maximum brightness levels, the amplitudes and the shapes of the light curve for many WTTS is evidence for migra- tion of the spots, within the limits of an active zone, and for changes in their size. It is interesting to note, that RS CVn and BY Dra stars show the same kind of long- term photometric behaviour and that active longitudes have long been reported on the Sun (Losh 1939). For RS CVn stars, active longitudes appear to persist for a number of years (Butler 1996). Thus, an activity mechanism with some similarities to that operating on the Sun may be indicated by these common observational characteris- tics with WTTS. Longer term observations have shown, that the phe- nomenon of stability of the initial epochs is most pro- nounced in seven WTTS: LkCa 4, LkCa 7, V410 Tau, V819 Tau, V827 Tau, V830 Tau and V836 Tau (Grankin 1997, 1998, 1999). Changes in the phase of minimum light for these stars did not exceed $\\pm 7\\%$ of the period on a time scale from 7 to 12 years. Similar results were obtained for 23 WTTS in the young Orion Nebula Cluster (Choi \\& Herbst 1996). Some interesting results have been received from anal\\-ysis of long-term light curves of 30 WTTS and 80 RS CVn and FK Com stars (Grankin et al. 2002). These authors have found that in 18 of the 110 stars monitored, an in- crease in the amplitude of the periodic variability accom- panies a rise of the maximum brightness level. This result is difficult to understand in the context of a simple model where the star contains only a single spotted region in a high latitude zone that is always visible, because increas- ing the size of the spot (as required to increase the am- plitude) should decrease the maximum brightness of the star. It is possible to explain such behavior by invoking a star covered with mutiple spots at different latitudes. In this case the amplitude will not depend just on the total area of the spots, but will also depend on the distribu- tion of these spots on the stellar surface. Similar results have been obtained recently for 24 WTTS stars in the ex- tremely young cluster IC 348 over five observing seasons (Cohen et al. 2004). It has been shown that the stability of average magnitude from season to season most likely indi- cates that spots on these WTTS stars tend to redistribute themselves and undergo changes in size, temperature, and location rather than simply appear or disappear {\\it en masse}. The observations to date support to the possible exis- tence of stellar activity cycles on WTTS. In an attempt to find and study such activity cycles we have contin- ued to monitor some tens of WTTS at Mt. Maidanak Observatory over many years. In this paper, we employ the unique and extensive Mt. Maidanak database to in- vestigate WTTS variability on a time scale of many years and, in some cases, multiple decades. In Section 2, we de- scribe the stellar sample and the observations carried out for the last 20 years at Mt. Maidanak. In Section 3, we dis- cuss the phenomenon of rotational modulation of the light curves of 36WTTS. In Section 4, we describe a spot model and some results of modeling. In Section 5, we define the statistical parameters used to characterize the long term variability of WTTS. Further modeling of the light curves will be presented in a later paper in this series. ", "conclusions": "Our main conclusion is that most WTTS have very sta- ble long term variability with relatively small changes of amplitude or mean light level. The long term variability seen merely reflects modulation in the cold spot distribu-tions. Photometric periods are stable over many years, as is the phase of minimum light for most objects, within some range. On the long term, the spot properties do change in subtle ways, leading to some variations in the shape and amplitudes of the light curves. The long term variability of WTTS is thus perhaps best understood by assuming that they are covered by a number of spots un- evenly distributed on the stellar surface and located at preferential longitudes and/or latitudes, much like active belts in the Sun and RS CVn systems, rather than result- ing from a single polar spot. Such an uneven spot distri- bution would account for the reported correlation between amplitude and brightness level, as well as for the quasi- sinusoidal shape of the light curves, which indicates that the unspotted photosphere is not seen in these stars at any rotational phase. It also implies that the coverage of the stellar photosphere by cold spots may be much larger than anticipated from simple modelling of the light curves, which only provide lower limits to the spot areal coverage. Doppler imaging of a sample of typical WTTS could be an effective means to obtain further constraints on the distribution of cold spots at the surface of WTTS. Modelling of the light curves for V819 Tau, V827 Tau, V830 Tau and V836 Tau indicates that the spotted regions cover as much as 30 to 90\\% of the visible stellar hemisphere and that the mean temperature of the spots is 500--1400~K lower than the ambient photosphere. More detailed model calculations of the light curve of V410 Tau showed that: (i) The decrease in mean brightness is attributable to the increase in total spot area from 47 to 53\\% and that it is essentially independent of the degree of nonuniformity in the spot distribution over the surface. (ii) The amplitude of the phased light curve depends on the degree of nonuniformity in the spot distribution more strongly than on the star\u0092s total spot area. An in- crease in amplitude was accompanied by an increase in the degree of nonuniformity in the spot distribution over the stellar surface from 21 to 35\\%. (iii) There is a weak correlation between the star's to- tal spot area and its degree of nonuniformity in the spot distribution with the correlation coefficient k=0.71. The increase in the star\u0092s total spot area from 47 to 53\\% was accompanied by a decrease in the degree of nonuniformity in the spot distribution from 35 to 21\\%. The most active variable stars in our sample (i.e. those with the largest amplitudes) demonstrate stability of phase of minimum light most strongly over many years. The data presented here should prove useful in fur- ther investigations of the nature and evolution of spots on WTTS." }, "0801/0801.2005_arXiv.txt": { "abstract": "We present a possible star formation and chemical evolutionary history for two early-type galaxies NGC~1407 and NGC~1400. They are the two brightest galaxies of the NGC~1407 (or Eridanus-A) group, one of the 60 groups studied as part of the Group Evolution Multi-wavelength Study (GEMS). Our analysis is based on new high signal-to-noise spatially resolved integrated spectra obtained at the ESO 3.6m telescope, out to $\\sim$~0.6 (NGC~1407) and $\\sim$~1.3 (NGC~1400) effective radii. Using Lick/IDS indices we estimate luminosity-weighted ages, metallicities and $\\alpha$-element abundance ratios. Colour radial distributions from HST/ACS and Subaru Suprime-Cam multi-band wide-field imaging are compared to colours predicted from spectroscopically determinated ages and metallicities using single stellar population models. The galaxies formed over half of their mass in a single short-lived burst of star formation ($\\geq$~100~M$_{\\sun}$/year) at redshift z~$\\geq$~5. This likely involved an outside-in mechanism with supernova-driven galactic winds, as suggested by the flatness of the $\\alpha$-element radial profiles and the strong negative metallicity gradients. Our results support the predictions of the revised version of the monolithic collapse model for galaxy formation and evolution. We speculate that, since formation the galaxies have evolved quiescently and that we are witnessing the first infall of NGC~1400 in the group. ", "introduction": "The physical formation and evolution processes taking place at various locations within a galaxy leave their fossil imprint in the stars. These chemodynamical imprints are the regular features observed in the stellar populations, photometry, internal kinematics and structure of early-type galaxies. Furthermore, these features are found to correlate defining scaling relations of remarkable tightness, such as the colour-magnitude relation and the fundamental plane. This class of galaxy, therefore, is one of the best guides to compare observations and theoretical model predictions. Theoretical scenarios that describe the formation and evolution of early-type galaxies include: a revised version of monolithic collapse, the dissipative (wet/gas-rich) and dissipationless (dry/gas-poor) merger alternatives of hierarchical clustering. In the classical models of monolithic collapse (\\citealt{eggen62}; \\citealt{larson74a}; \\citealt{larson75}; \\citealt{carlberg84}; \\citealt{arimoto87}), stars form in all regions during a rapid collapse and remain in their orbits, whereas the gas dissipates to the centre of the galaxy. The shape and ellipticity of stellar orbits are expected to depend on the initial angular momentum of the collapsing protogalactic gas cloud and on the effect of gas viscosity in redistributing angular momentum, towards outer radii, during the collapse. The sinking gas is continuously enriched by evolving stars. As a consequence, stars formed in the centre are predicted to be more metal-rich than those born in the outer galaxy regions. The efficiency of this process is proportional to the depth of the galaxy potential well, so that a strong correlation between metallicity gradient and galaxy mass is expected. The evolving stars are also responsible for the $\\alpha$-element enrichment of the interstellar medium (ISM). To reproduce the supersolar abundance of $\\alpha$-elements with respect to the iron-peak elements, observed in early-type galaxies, the models require that the collapse and star forming process occurred on timescales $\\la$~1~Gyr, which will produce null or very small age gradients (\\citealt{arimoto87}; \\citealt{matteucci94}; \\citealt{thomas99}). Recently, feedback processes such as supernova-driven galactic winds have been re-considered in more detailed numerical models of dissipative collapse (\\citealt{larson74b}; \\citealt{arimoto87}; \\citealt{gibson97}; \\citealt{martinelli98}; \\citealt{chiosi02}; \\citealt{kawata03}; \\citealt{pipino06}; \\citealt{pipino07}). It is suggested that they may create and shape the abundance gradients. Galactic winds initiate when the energy injected by supernovae explosion balances the gravitational binding energy of the ISM. The winds evacuate the gas, eliminating the essential fuel for any future star formation. The shallower local potential well of the external galactic regions supports the early development of winds with respect to the central regions. Therefore, in the central regions the star formation and the chemical enrichment last longer. Dissipation of gas towards the galaxy centre and a time-delay in the occurrence of galactic winds cooperate in steepening any metallicity gradient. In summary, dissipative collapse models predict: i) very steep negative metallicity gradients, that strongly correlate with galaxy mass (\\citealt{chiosi02}; \\citealt{kawata03}), ii) small or null age gradients, as a consequence of the short timescales involved in the collapse and star formation processes, iii) $\\alpha$-element enhancement gradients that can be either positive, negative or null (\\citealt{martinelli98}; \\citealt{pipino06}), due to the ``\\textit{interplay between local differences in the star formation timescale and gas flows}\" (\\citealt{pipino07}), iv) isophotal shape and ellipticity proportional to the protogalactic initial angular momentum and its radial redistribution, v) no peculiar internal kinematic structure. The difference between the original and revised dissipative collapse models focuses on the assembly of the initial gaseous material. The revised view accounts for the possibility of either a unique primordial cloud or the coalescence of many gaseous clumps without any preexisting stars. Otherwise, the physical processes in galaxy formation are similar to that described above. In the hierarchical clustering scenario of galaxy formation (eg. \\citealt{kauffmann93}; \\citealt{cole94}; \\citealt{baugh96}; \\citealt{kauffmann98}; \\citealt{delucia06}), elliptical galaxies are produced by merging events. Early numerical simulations of galaxy mergers predicted contradictory radial variations of stellar properties in merger remnants. \\cite{white80} suggested that dissipationless mergers cause a flattening of metallicity gradients, whereas \\cite{albada82} argued that the position of the stars in the local potential are preserved by violent relaxation and the gradients in the progenitors are only affected moderately. Further simulations of \\cite{barnes91}, considered the hydrodynamical physics of the gas and found that during a dissipative merger a significant fraction of the progenitors' gas tends to migrate toward the central regions of the merger remnant. \\cite{mihos94} showed that the gas accumulated in the central regions may trigger a local secondary burst of star formation. Galaxy merger-induced star formation is predicted to produce metallicity and age gradients. Positive or negative $\\alpha$-element enhancement gradients may also originate, depending on the original enhancement of the gas and the duration of the burst (\\citealt{thomasgre99}; \\citealt{thokauf99}). \\cite{bekki99} simulated dissipative mergers of two gas-rich disc galaxies of varying mass to form an elliptical galaxy remnant. Their models show that metallicity gradients in merger remnants are shallower with respect to those predicted by dissipative collapse models and only weakly dependent by the remnant galaxy mass. \\cite{kobayashi04} simulated the formation and chemodynamical evolution of elliptical galaxies in a cold dark matter cosmology, focusing on internal metallicity gradients. The models ranged from a monolithic collapse scenario, seen as the assembly of tens of gas-rich subunits at high redshift, to dissipative and dissipationless mergers of equal-mass galaxies at low redshift (which she denoted ``major mergers''). The result was that galaxies of a given mass had steep metallicity gradients if formed by collapse and shallower gradients if formed by major merger (dissipative or dissipationless). Recently, combining the Millennium cosmological N-body simulation (\\citealt{springel05}) with semi-analytic models, \\cite{delucia06} and \\cite{delucia07} focused on the distinction between the formation and assembly time (redshift) of elliptical galaxies and brightest cluster galaxies (BCGs) in the hierarchical clustering scenario. The models make use of the prescription of feedback from a central active galactic nucleus (AGN) and supernova (see \\citealt{croton06}) to prevent excessive galaxy-mass growth from cooling flows. Their model predicts a mass-dependent evolutionary history (``down sizing'' scenario), with more massive galaxies (e.g BCGs, BGGs; brightest group galaxies) forming a large fraction of their stars (50~$-$~80 per cent) in progenitor systems at redshift $\\sim$~2.5. The progenitors then assemble into a single final object at z~$\\sim$~0.8. The formation epoch of less massive galaxies is predicted to be z~$\\sim$~1.9 and the assembly time at redshift $\\sim$~1.5. The peak of the star formation history of massive galaxies is at redshift $\\sim$~5, and it progressively moves towards lower redshifts for less massive systems. Similarly, the duration of the star forming episode is predicted to last longer in low mass galaxies. Mergers are thought to induce bursts of star formation. Therefore, the star formation history of a galaxy presents the ``bursty'' behaviour described in \\cite{delucia06}. The kinematics and internal structure of a merger remnant are a consequence of the progenitors' mass ratios and the efficiency of the dissipative process (e.g. \\citealt{naab06}). \\cite{naab03} show that a dissipationless binary merger of equal-mass disk galaxies leads to a slowly rotating and anisotropic supported system ($(v/\\sigma)^{*}<$~0.4) with preferentially boxy isophotes and significant minor-axis rotation. Whereas, the merger of unequal-mass disk galaxies produces a rotationally supported elliptical galaxy with a small amount of minor-axis rotation ($(v/\\sigma)^{*}=$~1.2; suggesting isotropic oblate rotators) and disky isophotes. In dissipative mergers the gas is found to settle at the galaxy centre deepening the potential well and making it more axisymmetric. The isophotal shape of an equal-mass merger is then strongly affected by the gas, since most of the stars are gravitationally induced to move from boxy to more elliptical or even disky orbits. Similarly, the disky isophotal shape of an unequal-mass merger remnant is weakly affected by the presence of gas (\\citealt{naab06}). Furthermore, in simulations of an equal-mass merger with induced star formation (\\citealt{bekki97}) gradual gas dissipation forms an elliptical galaxy with disky isophotes, whereas a merger with rapid star formation produces both boxy and disky isophotes depending on the viewing angle of the observer. The merger of unequal-mass gas-rich disk galaxies is found to exhaust a large amount of the gas owing to moderate star formation and ending in the formation of an S0 galaxy (\\citealt{bekki98}). In summary, hierarchical galaxy formation models predict: i) shallow negative metallicity gradients, with little host mass dependence, ii) age gradients and iii) $\\alpha$-element enhancement gradients, that can be either positive or negative, due to the duration and location of the merger-induced star formation and the original abundance pattern of the gas, iv) isophotal shape and ellipticity depending on the progenitors mass ratio and the efficiency of gas dissipation, v) peculiar internal kinematic structures. The contrasting predictions of the galaxy formation models underline the lack of complete understanding of the physical processes behind the formation of galaxies. Observationally, we now have the ability to obtain spatially resolved high signal-to-noise integrated spectra out to large galactic radii (and hence larger mass fractions). This allows us to consider the physical mechanisms acting locally and how their properties vary with radius (i.e. investigating beyond the central regions), proving invaluable constraints on the processes of galaxy formation and evolution. Recently, an increasing number of works have focused their attention on stellar population radial profiles of galaxies in different environments (e.g. \\citealt{kobaari99}; \\citealt{mehlert03}; \\citealt{sanchez06}; \\citealt{sanchez07}; \\citealt{brough07}; \\citealt{reda07}). The results underline a common behaviour for the stellar population radial profiles of the studied galaxies: early-type galaxies are often characterised by strong metallicity gradients, shallow age gradients and null or statistically insignificant $\\alpha$-element enhancement gradients. Following from \\cite{spola08a} (hereafter Paper~I) we probe a possible star formation and chemical evolutionary history of the early-type galaxies NGC~1407 and NGC~1400. We accomplish the task using the spectral Lick/IDS indices (\\citealt{faber85}; \\citealt{worthey94}) to determine the radial profiles in age, metallicity, [Z/H], and $\\alpha$-element abundance ratios, [$\\alpha$/Fe], out to $\\sim$~0.6 (NGC~1407) and $\\sim$~1.30 (NGC 1400) times the galaxies' effective radii ($r_{e}$; the radius within which half the galaxy light is contained). Central values and gradients in age, [Z/H] and [$\\alpha$/Fe] are estimated and used as proxies of the quantity, velocity and duration of gas dissipation, star formation and possible merger events during galaxy formation. Furthermore, we have considered the results from the spatially resolved radial kinematics and surface photometry analysis performed in Paper~I. We compared colour radial profiles with colours predicted from spectroscopically derived ages and metallicities using single stellar population models. This paper is organised as follows. In Section~2 we describe the two sample galaxies. In Sections~3 and 4 we describe the observations and the relevant data reduction. Section~4 presents the spatially resolved stellar population parameters. In Section~5 observed and predicted colour index radial profiles are examined. In Section~6 we summarise our results, and discuss the possible star formation and evolutionary histories of NGC~1407 and NGC~1400. In Section~7 conclusions are presented. ", "conclusions": "We find that the stellar population of NGC~1407 is uniformly old, 12.0~$\\pm$~1.1~Gyr, with a supersolar degree of $\\alpha$-elements enhancement, 0.41~$\\pm$~0.05. We also measure a steep negative metallicity gradient of $-$0.38~$\\pm$~0.04 and a central metallicity of 0.29~$\\pm$0.08. The stellar population of NGC~1400 is found to be constantly old, 13.8~$\\pm$~1.1 Gyr, and largely enriched by $\\alpha$-elements, 0.33~$\\pm$~0.04. The central metallicity is 0.25~$\\pm$~0.06 and it steeply declines towards outer radii with a gradient of $-$0.47~$\\pm$~0.04. The results of this work (and Paper~I) suggest a similar star formation and evolutionary history for the two galaxies. The outlined scenario is compatible with the revised version of the monolithic collapse model of galaxy formation and evolution. The two galaxies formed the bulk of their stars at z~$\\geq$~5. The star forming episode lasted for no longer than 1~Gyr with a star formation rate $\\geq$~100~M$_{\\sun}$/year. NGC~1407 might have experienced an ancient merger, as inferred by the possible kinematically decoupled core (Paper~I); nevertheless, the detection is uncertain and potentially originated by a misalignment of the slit with respect to the centre of the galaxy, as suggested by the $h_{3}$ radial profile. In Paper~I we have also claimed that NGC~1400 has not interacted with NGC~1407 or the group intergalactic medium. The flatness of the $\\alpha$-element radial profiles and the strong negative metallicity gradients are consistent with an outside-in formation scenario, where supernova-driven galactic winds played a fundamental role and shaped the slope of the gradients. The uniform radial profile of old ages, steep metallicity gradients and lack of signature of recent mergers indicate that the galaxies formed stars quickly some $\\sim$~12~Gry ago and then have evolved quiescently without any significant interaction ever since." }, "0801/0801.0693_arXiv.txt": { "abstract": "We present the strategies adopted in the relative and absolute calibration of two different data sets: $U,B,V,I$-band images collected with the Wide Field Imager (WFI) mosaic camera mounted on the 2.2m ESO/MPI Telescope and $u,v,b,y$ \\calastrom images collected with the 1.54m Danish Telescope (ESO, La Silla). In the case of the WFI camera we adopted two methods for the calibration, one for images collected before 2002, with the ESO filters $U/38_{ESO841}$ and $B/99_{ESO842}$, and a different one for data secured after 2002, with the filters $U/50_{ESO877}$ and $B/123_{ESO878}$. The positional and color effects turned out to be stronger for images collected with the old filters. The eight WFI chips of these images were corrected one by one, while in the case of images secured with the new filters, we corrected the entire mosaic in a single step. In the case of the Danish data set, we compared point-spread function (PSF) and aperture photometry for each frame, finding a trend in both the X and Y directions of the chip. The corrections resulted in a set of first and second order polynomials to be applied to the instrumental magnitudes of each individual frame as a function of the star position. ", "introduction": "\\subsection{Observations and data reduction} Data have been retrieved from the ESO archive and include 8 $U$, 39 $B$, 51 $V$, and 26 $I$ images of \\calaomc. Data include both shallow and relatively deep images, with exposures times ranging from 1 to 300s for the $B, V, I$ bands, and from 300 to 2400s for the $U$ band, and were collected in several observing runs ranging from 1999 to 2003. During this period two filters were changed: data secured before 2002 were collected with the filters $U/38_{ESO841}$ and $B/99_{ESO842}$, while later ones with the filters $U/50_{ESO877}$ and $B/123_{ESO878}$. These data were obtained in good seeing conditions, and indeed the mean seeing ranges from ~0.6\" for the $I$ band to ~1.1\" for the $U$ band. We accurately selected the best PSF stars uniformly distributed across each chip. A moffat analytical function linearly variable on the chip was assumed for the PSF. The data were reduced with DAOPHOT {\\small IV}/ALLFRAME (Stetson 1994). \\subsection{Relative and absolute calibration} We are interested in providing accurate relative calibration of individual chips of the WFI mosaic camera because the occurrence of positional effects might cause a spurious color broadening of key evolutionary features such as the Red-Giant Branch (RGB) and the Main-Sequence Turn-Off. Moreover, we need to provide an accurate absolute calibration of our data in order to compare theory with observations. The plausibility of this comparison relies on the accuracy of the absolute zero-point of individual bands. This is a fundamental requirement for the distance modulus, and in turn for the absolute ages. Recent findings (Corsi et al. 2003; Koch et al. 2004) based on photometric data collected with the WFI indicate that the eight CCD chips might be affected by positional effects involving zero-point errors of the order of several hundredths of magnitude. This subtle effect could be due to scattered light. It seems that telescopes equipped with mosaic cameras and focal reducers may present this problem (Manfroid et al. 2001, Manfroid \\& Selman 2001). \\begin{floatingfigure}[l]{0.45\\textwidth} \\centering \\includegraphics[height=5.cm,width=5.5cm]{calamidaF1cut.ps} \\caption{Residuals, $\\Delta mag = mag_S - mag_{PSF}$, in the $V$ band, plotted versus the X, Y position on the chip, before (top) and after (bottom) the positional corrections were applied, for an image collected in 2002.} \\end{floatingfigure} In order to correct the positional effects of the WFI mosaic camera and to perform an accurate absolute calibration of \\calaomc data set, we followed these steps:\\\\ $\\bullet$ obtain a set of local standard stars for \\calaomc;\\\\ $\\bullet$ identify all the trends of PSF magnitudes compared to standard magnitudes versus position on the frame and correct them;\\\\ $\\bullet$ estimate the calibration curves.\\\\ We thus made use of a set of new multi-band ($U,B,V,I$) local standard stars for \\calaomc (Stetson et al. 2007). This star list has been selected in photometric accuracy ($\\sigma \\leq$ 0.03 mag) and in 'separation index' ($sep \\geq$ 2.5). We ended up with a catalog of $\\sim 3\\times 10^4$ local standard stars. The sky area covered by these stars is $\\sim$ 37'$\\times$40', and includes a substantial fraction of our WFI data. We thus applied two methods: in the case of images collected before 2002 ($U/38_{ESO841}$ and $B/99_{ESO842}$ filters), the eight chips were corrected one by one, while for images secured starting from 2002 ($U/50_{ESO877}$ and $B/123_{ESO878}$ filters) we corrected the entire mosaic in a single step. The color terms for these filters have a very steep slope, therefore, we decided to estimate a first color curve and apply it. We then studied the residuals, $\\Delta mag = mag_S - mag_{PSF}$, where $S$ stands for Standard, as a function of the X and Y position on the chip (or mosaic, see Fig. 1). Once corrected for the positional effects, we estimated the color term once again and applied it for the absolute zero-point calibration. This strategy relies on the assumption that the two corrections are independent, and there is no reason why this should not be the case. Fig. 1 shows the magnitude residuals in the $V$ band plotted versus the X and Y positions on the frame, before and after the corrections were applied to an image collected in 2002. The trend with the position is clear, non-linear, and stronger in the four outermost chips. A small residual trend on the position is still present in the external regions, where the lack of standard stars makes it difficult to estimate the corrections. \\begin{floatingfigure}[l]{0.45\\textwidth} \\centering \\includegraphics[height=5cm,width=5cm]{calamidaF2cut.ps} \\caption{Residuals, $\\Delta mag = mag_{AP} - mag_{PSF}$, in the $u$ band, plotted versus the X, Y position on the chip, before (top) and after (bottom) the positional corrections were applied.} \\end{floatingfigure} We thus corrected the positional effects by fitting the $\\Delta mag$ vs $X/Y$ with second, third or fourth order polynomials, and we then estimated the calibration curves for each set. We adopted a first order polynomial to calibrate the $V$ and the $I$ bands as a function of the instrumental $V-I$ color. In the case of the $B$ band, we used a first order polynomial to calibrate the new filter, $B/123_{ESO878}$, and a third order one for the old filter, $B/99_{ESO842}$, as a function of the instrumental $B-V$ color. Particular attention has been paid to the $U$ band calibration, either in the case of the old filter, $U/38_{ESO841}$, as well in the case of the new one, $U/50_{ESO877}$. Having applied the positional corrections to each set, we estimated a first color curve, a fourth order polynomial for $U$ vs $U-I$. After applying this calibration curve, the $U$ magnitudes still showed a residual trend as a function of the $U$ magnitude and the $U-I$ color. We thus corrected these trends with first and sixth order polynomials, respectively. The reason of these very complicated trends with colors is probably due to the shape of the old $B$ and $U$ filters, which is quite different from the shape of the standard Johnson filters. ", "conclusions": "" }, "0801/0801.2375_arXiv.txt": { "abstract": "We present integrated $JHK_s$ 2MASS photometry and a compilation of integrated-light optical photoelectric measurements for 84 star clusters in the Magellanic Clouds. These clusters range in age from $\\approx200$~Myr to $>10$~Gyr, and have [Fe/H] values from $-2.2$ to $-0.1$ dex. We find a spread in the intrinsic colours of clusters with similar ages and metallicities, at least some of which is due to stochastic fluctuations in the number of bright stars residing in low-mass clusters. We use 54 clusters with the most reliable age and metallicity estimates as test particles to evaluate the performance of four widely used SSP models in the optical/NIR colour-colour space. All models reproduce the reddening-corrected colours of the old ($\\ge$ 10 Gyr) globular clusters quite well, but model performance varies at younger ages. In order to account for the effects of stochastic fluctuations in individual clusters, we provide composite $B-V$, $B-J$, $V-J$, $V-K_s$ and $J-K_s$ colours for Magellanic Cloud clusters in several different age intervals. The accumulated mass for most composite clusters are higher than that needed to keep luminosity variations due to stochastic fluctuations below the 10\\% level. The colours of the composite clusters are clearly distinct in optical-NIR colour-colour space for the following intervals of age: $>10$~Gyr, $2-9$~Gyr, $1-2$~Gyr, and $200$~Myr$-1$~Gyr. This suggests that a combination of optical plus NIR colours can be used to differentiate clusters of different age and metallicity. ", "introduction": "The most efficient method to determine the age and metallicity for unresolved stellar systems (especially at high redshift) is by comparing their observed colours with the predictions of evolution synthesis models \\citep[e.g.][]{bc93, bc03, worthey94, vazdekis99, maraston98, maraston05, af03}. Thus, it is important to test the integrated colours predicted by recent models, based on objects which have accurate ages and metallicities determined independently. In the present paper we focus our attention on the combination of visual and near-infrared (NIR) photometry, which has proven to be important for breaking the age-metallicity degeneracy, particularly in stellar populations older than $\\approx {\\rmn{a\\ few\\ times}} \\times 100$~Myr \\citep[e.g.][]{goudfrooij01, puzia02, hempel04}. With the advent of the {\\it Spitzer Space Telescope (Spitzer)\\/} and mid-infrared (MIR) instrumentation for some large ground-based telescopes, the NIR spectral region is now accessible at intermediate-to-high redshifts. In a recent paper based on {\\it Spitzer\\/} Infrared Array Camera (IRAC) imaging, \\cite{vanderwel2006} reported significant discrepancies between some model predictions and the observed rest-frame $K$-band properties of early-type galaxies at z $\\approx 1$. Their results show that the interpretation of NIR photometry is hampered by model uncertainties. As a consequence the determination of masses of distant stellar systems based on such data can have uncertainties up to a factor 2.5 \\citep[see][]{bruzual07}. Unfortunately, providing accurate model predictions in the near-infrared is challenging, since there are limitations imposed by the current lack of understanding of certain stages of stellar evolution (e.g., thermally pulsing asymptotic giant branch, or TP-AGB stars). These objects significantly affect the spectral energy distribution (SED) in the NIR and MIR for stellar populations with ages between $\\approx200$ Myr to 3 Gyr. Another possible complication is that the stellar libraries used by population synthesis models contain mostly stars from the solar neighborhood. These stars have a star formation history which is not necessarily typical for extragalactic populations (e.g. relatively little variation of [$\\alpha$/Fe] ratios), and there are only a very limited number of AGB spectra available. The Large and Small Magellanic Clouds (LMC and SMC respectively) provide a unique opportunity to test the accuracy of most current SSP models, since they contain a significant population of intermediate-age massive star clusters which are not easily accessible in our Galaxy. The ages and metallicities of these star clusters can be determined from deep colour-magnitude diagrams (CMDs) reaching below the main sequence turn-off (MSTO)\\footnote{Obtaining photometry with sufficient quality to secure reliable age and metallicity estimates for clusters in galaxies beyond the Magellanic Clouds requires a significant investment of observing time. To date only one such cluster, SKHB~312 in M31, has a CMD deep enough to probe the MSTO region \\citep{brown2004}. The photometry for this object was obtained as a result of a program utilizing 126 {\\it Hubble Space Telescope (HST)\\/} orbits. }. Medium and high-resolution spectroscopy of individual giants in these clusters also provides independent metallicity estimates. Therefore their integrated-light properties (easily observed with small and moderate-aperture telescopes) can be combined with the accurate age/metallicity measurements and used to test (and calibrate) the SSP models. In \\cite{pessev2006} (hereafter Paper~I) we used the Two Micron All Sky Survey (2MASS; \\cite{skrutskie2006}) to derive NIR $(JHK_s)$ integrated-light magnitudes and colours for a large sample of Magellanic Cloud star clusters, based on a homogeneous, accurately calibrated dataset. In the present study we use the sample from Paper~I and new photometry for 9 additional objects (forming the largest dataset of integrated NIR magnitudes and colours of LMC/SMC star clusters to date) to test the performance of several SSP models. We combine the 2MASS data with optical photometry originating from the work of \\cite{bica_et_al_96} and the compilation of \\cite{vdb81}. The technique adopted in Paper~I - measuring $JHK_s$ curves of growth to large radii allows us to utilize rather heterogeneous databases of optical photometry, usually performed with a set of fixed apertures. We use 54 clusters from our sample as ``test particles''. These clusters were chosen to have reliable age and metallicity measurements, covering a wide parameter space. \\setcounter{figure}{0} \\begin{figure} \\centering \\includegraphics[bb=14 14 256 256,width=8.15cm]{fig1.eps} \\caption{ A finding chart of the LMC showing the clusters in our sample. The $R$-band image is centred on $\\alpha_{2000} = 05^{h} 26^{m} 37.7^{s}$ and $\\delta_{2000} = -68^{o} 56' 57.5\"$. The arrow near the centre outlines the position of NGC~1928 ($\\alpha_{2000} = 05^{h} 20^{m} 57.7^{s}$ and $\\delta_{2000} = -69^{\\rmn{o}} 28' 40.2\"$). This cluster is located close to the geometrical centre of the LMC bar, and was adopted by Bica et al.\\ (1996) as a reference point for the relative coordinates of the LMC cluster system. The labeled arrows show the direction towards the clusters lying outside the boundaries of this chart. The $R$-band image (G. Bothun 1997, private communication) covers $8^{\\rmn{o}}$x$8^{\\rmn{o}}$, while the dimensions of the chart are $16^{\\rmn{o}}$x$16^{\\rmn{o}}$.} \\label{fig:lmc_map} \\end{figure} \\setcounter{figure}{1} \\begin{figure} \\centering \\includegraphics[bb=14 14 256 256,width=8.15cm]{fig2.eps} \\caption{A finding chart of the SMC showing the clusters in our sample. The $R$-band image (G. Bothun 1997, private communication) is centred on $\\alpha_{2000} = 01^{h} 04^{m} 42.8^{s}$ and $\\delta_{2000} = -72^{\\rmn{o}} 52' 32.4''$. SMC clusters cover a smaller area than LMC objects. There is only one cluster outside the $R$ frame, but for illustrational purposes the dimensions of this finder chart are identical to those of the LMC chart in Figure~\\ref{fig:lmc_map}.} \\label{fig:smc_map} \\end{figure} This paper is organized as follows: in \\S 2 we define our extended sample and present the new photometry along with the compilation of visual magnitudes and colours. Four sets of SSP models are tested in \\S 3, followed by concluding remarks in \\S 4. Information about the properties of the cluster sample is presented in Appendices A and B. Transformations between the model grids in the \\cite{bb88} system and the photometric system of 2 MASS are provided in Appendix C. \\label{intro} ", "conclusions": "We have presented new integrated $JHK_s$ 2MASS photometry for 9 Magellanic Cloud clusters, bringing our total sample (when combined with the results of Paper I) to 54 clusters with reliable ages $\\geq$200~Myr. In addition, we compile integrated-light $B$ and $V$ photometric measurements, extinction estimates, and a database of reliable age and metallicity determinations (mostly recent results) from the literature for our sample clusters. We divide the clusters into different age (e.g., $\\geq10$~Gyr, $3-9$~Gyr, $3-4$~Gyr, $1-2$~Gyr, and 200~Myr$-$1~Gyr) and metallicity (when possible), and quantify the observed spread in the intrinsic cluster colours in these ranges. Care was taken to account for the spread of the observational data around the model predictions due to the stochastic fluctuations in the stellar populations of the clusters. The smallest spread in intrinsic colours is found for clusters with ages $\\ga 10$~Gyr, the colours of which are well-reproduced by all four sets of SSP model predictions. The systematic shift between the model predictions and the observed colours for a sample of old Milky Way globular clusters reported by \\cite{cohen07} is not observed in our Magellanic Cloud cluster analysis. The largest spread in colour is found for clusters in the age range $2-4$~Gyr. We believe that much of the spread in the colours for {\\it individual\\/} clusters younger than 10~Gyr results from stochastic fluctuations in the numbers of infrared-luminous stars, since individual clusters tend to have less than ${\\cal{M}}_{\\rm (10\\%)}$\\footnote{stellar mass needed to decrease the the luminosity uncertainty due to stochastic effects in the stellar population to 10\\%)} contributing to the observed colours. Composite $(B-J)_0$, $(V-J)_0$, and $(J-K_s)_0$ cluster colours are calculated for each age/metallicity interval, and compared with the predictions of four widely used population synthesis models \\citep{maraston05, bc03, af03, vazdekis99}, in order to evaluate their performance. We interpolate the model grids to calculate the offset or distance in colour-colour space between the model predictions and the age and metallicity for our composite cluster colours. All four sets of models reproduce the colours of old ($\\geq10$~Gyr) Magellanic Cloud clusters quite well, with the \\citet{maraston05} and \\citet{bc03} models giving slightly better fits than the other two. In the age range of 2\\,--\\,10 Gyr, the \\citet{maraston05} models have the largest separation in optical-NIR colour-colour space between the 2~Gyr and 10~Gyr model tracks, which best reproduces our observed composite colours in the $2-3$~Gyr and $3-9$~Gyr ranges. While the composite colour for $2-3$~Gyr-old clusters falls just off the grid for the other three models, actual quantitative distances between the model predictions and composite cluster colours are not significantly different among the four models. In the $1-2$~Gyr and $0.2-1$~Gyr age ranges, the \\cite{bc03} models generally give the best quantitative match to our composite Magellanic Cloud cluster colours. Taking into account the inferred ages and metallicities, there is little difference between the \\cite{bc03} and \\cite{maraston05} model performance. The cluster colours fall off the \\citet{af03} and \\citet{vazdekis99} model predictions in the two youngest age ranges, largely due to their limited coverage at low metallicities. Based on the comparisons presented in this work, it is found that each model has strong and weak points when used to analyse the optical$+$NIR colours of unresolved stellar populations. There is no model set that clearly outperforms the others in all respects. Overall, the \\cite{bc03} and typically yield the best quantitative match to our composite cluster colors. The \\cite{maraston05} models are a close second. The same two models also yield the best match to the composite cluster ages and metallicities." }, "0801/0801.2972_arXiv.txt": { "abstract": "We present a sequence of toy models for irradiated planet atmospheres, in which the effects of geometry and energy redistribution are modelled self-consistently. We use separate but coupled grey atmosphere models to treat the ingoing stellar irradiation and outgoing planetary reradiation. We investigate how observed quantities such as full phase secondary eclipses and orbital phase curves depend on various important parameters, such as the depth at which irradiation is absorbed and the depth at which energy is redistributed. We also compare our results to the more detailed radiative transfer models in the literature, in order to understand how those map onto the toy model parameter space. Such an approach can prove complementary to more detailed calculations, in that they demonstrate, in a simple way, how the solutions change depending on where, and how, energy redistribution occurs. As an example of the value of such models, we demonstrate how energy redistribution and temperature equilibration at moderate optical depths can lead to temperature inversions in the planetary atmosphere, which may be of some relevance to recent observational findings. ", "introduction": "The past several years have seen a rapid and exciting increase in the amount of information available concerning the physical properties of extrasolar planets. In particular, the class of short orbital period planets, known informally as `Hot Jupiters', has begun to yield information through a variety of observational channels. The discovery of systems in which the orbital plane is almost edge-on to the line of sight has resulted in the detection of planetary transits (Charbonneau et al. 2000; Henry et al. 2000; Bouchy et al. 2005). Optical and ultraviolet photometry and spectroscopy of such systems can yield information about the absorbative properties of the planet atmosphere (Brown et al. 2001; Charbonneau et al. 2002; Vidal-Madjar et al. 2003; Vidal-Madjar et al. 2004). In addition, the unparalleled capabilities of the Spitzer Space Telescope has allowed the detection of thermal infrared emission (Charbonneau et al. 2005; Deming et al. 2005, Deming et al. 2006; Grillmair et al. 2007; Richardson et al. 2007) during secondary eclipses as well as the measurement of phase curves for non-transitting systems (Harrington et al. 2006). This flurry of observational activity has spurred a similar level of activity on the theoretical front, with several independent models being used to infer the physical properties of the planets from the observations (Seager \\& Sasselov 2000; Barman, Hauschildt \\& Allard 2001; Cho et al. 2003; Menou et al 2003; Sudarsky, Burrows \\& Hubeny 2003; Burrows, Hubeny \\& Sudarsky 2005; Dyudina et al. 2005; Cooper \\& Showman 2005; Seager et al. 2005; Barman, Hauschildt \\& Allard 2005; Fortney et al. 2005; Fortney 2005; Burrows, Sudarsky \\& Hubeny 2006; Cooper \\& Showman 2006; Fortney et al. 2006; Langton \\& Laughlin 2007). The resulting inferences are not always entirely in agreement. As an example, consider each group's response to the first detection of secondary eclipses. Burrows et al. (2005) concluded that the redistribution of energy from the substellar to the antistellar face of the planet was weak. On the other hand, Barman, Hauschildt \\& Allard (2005), as well as Fortney et al. (2005), concluded that the redistribution was strong, while the opinion of Seager et al. (2005) fell somewhere in between. Similarly, predictions for the effects of atmospheric fluid motions on the emergent intensity pattern differ substantially from group to group (Cho et al. 2003; Cooper \\& Showman 2006; Langton \\& Laughlin 2007). This diversity of opinion is hardly surprising, given the complexity of the problem. The study of Hot Jupiter atmospheres requires that we understand the atmospheric chemistry, radiative transfer and hydrodynamics of the planetary atmosphere. These are affected by the thermal structure and evolution of the planet, by the effects of photochemistry induced by ultraviolet light from the star, the possible existence and behaviour of atmospheric condensates, whether they form clouds, and whether (or if) energy is redistributed across the surface of the planet or reradiated in situ. Furthermore, there are a variety of different observational probes, each of which provides information but in ways that differ, sometimes subtly, from other methods. The complexity of the coupled radiative transfer, chemical equilibrium and atmospheric flow models also restrict somewhat the flexibility of the models in providing the true qualitative or geometric information necessary to develop a real understanding of what the observations are telling us. The overall intent in the pages to follow is pedagogical. Clearly, detailed model atmospheres are necessary to fully interpret the emerging observations. However, simplified models that focus on key physical elements can be useful to try and develop a qualitative understanding of what elements of the model each of the various extant observations is actually probing. To do this, we construct a sequence of analytic and semi-analytic toy models. These models are chosen to be simple enough to be transparent while still capturing the qualitative features of the more detailed models. The hope is that one can use this approach to tailor our interpretations of the more detailed calculations to reach a proper synthesis between theory and the different observations that we have at our disposal. In \\S~\\ref{Nada} we review our simplest model, based on a variation of the grey atmosphere model. In this first case we will assume no redistribution of energy across the face of the planet. In \\S~\\ref{Apply} we then examine how the various available observations relate to the underlying structure of the model. In \\S~\\ref{Finally} we then introduce the physics of energy redistribution into the model and examine how this affects the observations in \\S~\\ref{MoreApps}. Finally, in \\S~\\ref{Finis}, we place our models within the context of both observations and more detailed models in the literature, and examine possible implications for future generations of models. ", "conclusions": "In summary, the intent of the models presented here was to understand the qualitative behaviour of radiative transfer in hot Jupiter atmospheres. The simplified nature of the models is dictated by the difficulties of including both full radiative transfer and detailed hydrodynamics, both of which are important parts of a proper understanding of the phenomenon. In particular we have studied how the models change depending on the parameters $\\gamma$ (which reflects the difference in opacity for incoming optical-wavelength radiation and outgoing infrared-wavelength radiation) and $\\tau_w$ (the depth at which energy is redistributed horizontally across the planets atmosphere). We have also studied how the various new observational probes of Hot Jupiter atmospheres are influenced by changes in these parameters. We find that models in which redistribution occurs at large infra-red optical depths redistribute little energy to the night side (as expected). Perhaps more surprisingly, we find that significant day/night flux differences can also be found in the case $\\tau_w<1$, as long as $\\gamma < 1$ as well. We find that day-side temperature inversions are a generic feature of atmospheres in which the redistribution occurs at moderate ($\\tau_w \\sim 1$--1000) optical depths. This is of interest because of recent claims that temperature inversions have been found in HD~209458b (Knutson et al. 2007). The initial explanation for such behaviour is high altitude absorption (Burrows et al. 2007; Fortney et al. 2007), although the very low albedo (Rowe et al. 2007) restricts the absorbing species to be a poor reflector. Some level of redistribution is required to explain the HD~209458~b secondary eclipse amplitude and (lack of) phase variations, so that another potential explanation for the temperature inversion may simply be that it is related to the depth at which the energy is removed from the day-side. Certainly, figures~\\ref{BH6} and \\ref{BH7} show that some parts of the planet experience pronounced temperature inversions, even without a significant high altitude absorber. However, it will require detailed radiative transfer models to confirm whether this is a viable model or not. Our initial exploration has been far from exhaustive. One important advantage of these toy grey atmosphere models is their flexibility, and we have studied how energy redistribution changes the character of the models. This will hopefully provide a guide as to how one might perform similar calculations using the more detailed models, which presently can only consider changes at either the top or bottom of the atmosphere. Furthermore, we have also not exhausted the parameter space of toy models. For instance, one can imagine many other redistribution schemes apart from the one we have implemented here." }, "0801/0801.0977_arXiv.txt": { "abstract": "An excess over the extrapolation to the extreme ultraviolet and soft X-ray ranges of the thermal emission from the hot intracluster medium has been detected in a number of clusters of galaxies. We briefly present each of the satellites (EUVE, ROSAT PSPC and BeppoSAX, and presently XMM-Newton, Chandra and Suzaku) and their corresponding instrumental issues, which are responsible for the fact that this soft excess remains controversial in a number of cases. We then review the evidence for this soft X-ray excess and discuss the possible mechanisms (thermal and non-thermal) which could be responsible for this emission. ", "introduction": "\\label{Introduction} The existence of soft excess emission originating from clusters of galaxies, defined as emission detected below 1~keV as an excess over the usual thermal emission from hot intracluster gas (hereafter the ICM) has been claimed since 1996. Soft excesses are particularly important to detect because they may (at least partly) be due to thermal emission from the Warm-Hot Intergalactic Medium, where as much as half of the baryons of the Universe could be. They are therefore of fundamental cosmological importance. Soft excess emission has been observed (and has also given rise to controversy) in a number of clusters, mainly raising the following questions: 1)~Do clusters really show a soft excess? 2)~If so, from what spatial region(s) of the cluster does the soft excess originate? 3)~Is this excess emission thermal, originating from warm-hot intergalactic gas (at temperatures of $\\sim 10^6$~K), or non-thermal, in which case several emission mechanisms have been proposed. Interestingly, some of the non-thermal mechanisms suggested to account for soft excess emission can also explain the hard X-ray emission detected in some clusters, for example by RXTE and BeppoSAX (also see \\citealt{petrosian2008} - Chapter 10, this volume; \\citealt{rephaeli2008} - Chapter 5, this volume). Several instruments have been used to search for soft excess emission: EUVE, ROSAT and BeppoSAX in the 1990's, and presently XMM-Newton, Chandra and Suzaku. We will briefly present a history of these detections, emphasizing the difficulties to extract such weak signal and the underlying hypotheses, and summarising what is known on each of the observed clusters. For clarity, results will be presented separately for the various satellites. Finally we summarise and discuss the various models proposed to account for Extreme Ultraviolet (hereafter EUV) emission with their pros and cons. ", "conclusions": "\\label{conclusions} While in most cases the origin of the soft excess emission is difficult to prove unambiguously, due to problems with instrument calibration and unknown background/foreground emission level, the independent detection of the phenomenon with many instruments gives confidence in the genuine nature of the phenomenon in a number of clusters. Both thermal and non-thermal emission mechanisms are probably at work in producing the soft excess emission in clusters." }, "0801/0801.3659_arXiv.txt": { "abstract": "Since type Ia Supernovae (SNe) explode in galaxies, they can, in principle, be used as the same tracer of the large-scale structure as their hosts to measure baryon acoustic oscillations (BAOs). To realize this, one must obtain a dense integrated sampling of SNe over a large fraction of the sky, which may only be achievable photometrically with future projects such as the Large Synoptic Survey Telescope. The advantage of SN BAOs is that SNe have more uniform luminosities and more accurate photometric redshifts than galaxies, but the disadvantage is that they are transitory and hard to obtain in large number at high redshift. We find that a half-sky photometric SN survey to redshift $z = 0.8$ is able to measure the baryon signature in the SN spatial power spectrum. Although dark energy constraints from SN BAOs are weak, they can significantly improve the results from SN luminosity distances of the same data, and the combination of the two is no longer sensitive to cosmic microwave background priors. ", "introduction": "Type Ia supernovae\\footnote{We consider only type Ia SNe in this \\emph{Letter}.} (SNe) have become a mature tool for studying the cosmic expansion history \\citep[e.g.,][]{phillips93,riess98, perlmutter99a}. A number of SN surveys, such as the Sloan Digital Sky Survey (SDSS) II \\citep{frieman08}, the Supernova Legacy Survey \\citep{astier06}, and the ESSENCE Supernova Survey \\citep{miknaitis07} are being carried out to improve the statistics and our understanding of the systematics. Moreover, the SN technique will be an integral part of almost every proposed dark energy survey including the Large Synoptic Survey Telescope\\footnote{See \\url{http://www.lsst.org}.} \\citep[LSST, see][]{tyson06} and the Joint Dark Energy Mission. The conventional SN technique, measuring only the relative luminosity distance, $D_{\\rm L}$, is subject to degeneracies between cosmological parameters. For example, the SN constraint on the dark energy equation-of-state (EOS, $w$) parameter $w_{\\rm a}$, as defined by $w(z) = w_0 + w_{\\rm a}z/(1+z)$, is sensitive to the prior on the mean curvature of the universe \\citep*{linder05b, knox06c}. The reason is that the response of the relative distance to a variation in $w_{\\rm a}$ resembles that to a variation in the mean curvature and that the SN technique lacks the calibration of absolute distances \\citep{zhan06e}. Even for a flat universe with $w(z)\\equiv w_0$, the SN constraint on $w_0$ can be tightened considerably if the matter density is known to high precision \\citep{frieman03}. Such priors may come from other techniques, such as the cosmic microwave background (CMB), weak lensing, and baryon acoustic oscillations \\citep*[BAOs,][]{eisenstein98, cooray01b, blake03, hu03b, linder03b, seo03}. It has indeed been demonstrated that the latter three techniques are highly complementary to the SN $D_{\\rm L}$ technique \\citep{frieman03, seo03, knox06c}. Since SNe explode in galaxies, their distribution bears the BAO imprint as well. To measure the SN spatial power spectrum, one needs the angular position and redshift of each SN, not its luminosity. Hence, the SN BAO technique does not suffer from uncertainties in the SN standard candle, which constitute the largest unknown in the $D_{\\rm L}$ measurements. Nevertheless, the narrow range of the SN intrinsic luminosity reduces the effect of Malmquist-like biases and luminosity evolution. The SN rate traces the mass and star formation of the host galaxies with a time delay \\citep{sullivan06}. This means that SNe have a different clustering bias than galaxies that are selected by their luminosity or color. Finally, SNe have rich and time-varying spectral features for accurate estimation of photometric redshifts (\\phz{}s) \\citep*{pinto04, wang07a, wang07b}, which is helpful for measuring BAOs from a photometric survey. There have been discussions of using the SN weak lensing magnification \\citep*{cooray06} and nearby SN peculiar velocities \\citep*{hannestad07} to probe the large-scale structure. We focus on photometric SN surveys for BAOs in this \\emph{Letter} and note in passing that the SN weak lensing technique is more limited by shot noise than the SN BAO technique \\citep{zhan06d, zhan06e} and that the SN peculiar velocity technique requires precise redshift and distance measurements. For the BAO technique to be useful, one must survey a large volume at a sufficient sampling density as uniformly as possible. Although SN events are rare, the spatial density of SNe accumulated over several years will be comparable to the densities targeted for future spectroscopic galaxy BAO surveys. ", "conclusions": "We have demonstrated that \\phz{} SN data can be used to measure BAOs and to constrain cosmological parameters. The BAO constraints on the matter density and the baryon density are sensitive to the priors on the curvature and the scalar spectral index but not to the dark energy parameters. The dark energy constraints from the SN BAO technique alone are not meaningful. However, a combination of the SN BAO and $D_{\\rm L}$ techniques reduces the EP considerably, and the \\emph{Planck} priors are no longer crucial. The SN BAO results in Section~\\ref{sec:cons} are also applicable to \\phz{} galaxy BAOs. We note that \\phz{} errors are a large uncertainty for \\phz{} SN cosmology. Although our assumption about them is conservative compared to the results in \\citet{pinto04}, further studies are needed to make realistic forecasts. Long and non-uniform cadence may result in uneven sampling of the SN spatial distribution. Fortunately, LSST will be likely to always catch the SNe (especially high-$z$ ones) at their maximum owing to its fast sky coverage and rapid sampling. Furthermore, the effect of the cadence on the SN depth can be simulated and determined, and methods of correcting for uneven depths in galaxy surveys can be applied to the SN data. A spectroscopic SN BAO survey will be impractical, because one would have to revisit the sky many times spectroscopically over thousands of square degrees to catch the SNe that occur at different times. However, LSST will be able to obtain SNe in the millions over half the sky photometrically. This opens a window for applying the BAO technique to SNe and achieving more robust constraints with \\phz{} SN data. Since this SN BAO analysis requires no additional observations than doing the SN $D_{\\rm L}$ analysis alone, it should be a feature of all large-area SN cosmology analyses. Moreover, the SN data can also help calibrate the host galaxy \\phz{}s and the \\phz{} error distribution of other galaxies through the cross-correlation method \\citep{schneider06,zhan06d,newman08}." }, "0801/0801.4286_arXiv.txt": { "abstract": "We present results from {\\it Hubble Space Telescope} ultraviolet spectroscopy of the massive X-ray and black hole binary system, HD~226868 = Cyg X-1. The spectra were obtained at both orbital conjunction phases in two separate runs in 2002 and 2003 when the system was in the X-ray high/soft state. The UV stellar wind lines suffer large reductions in absorption strength when the black hole is in the foreground due to the X-ray ionization of the wind ions. A comparison of the {\\it HST} spectra with archival, low resolution spectra from the {\\it International Ultraviolet Explorer Satellite} shows that similar photoionization effects occur in both the X-ray high/soft and low/hard states. We constructed model UV wind line profiles assuming that X-ray ionization occurs everywhere in the wind except the zone where the supergiant blocks the X-ray flux. The good match between the observed and model profiles indicates that the wind ionization extends to near to the hemisphere of the supergiant facing the X-ray source. We also present contemporaneous spectroscopy of the H$\\alpha$ emission that forms in the high density gas at the base of the supergiant's wind and the \\ion{He}{2} $\\lambda4686$ emission that originates in the dense, focused wind gas between the stars. The H$\\alpha$ emission strength is generally lower in the high/soft state compared to the low/hard state, but the \\ion{He}{2} $\\lambda4686$ emission is relatively constant between X-ray states. The results suggest that mass transfer in Cyg~X-1 is dominated by the focused wind flow that peaks along the axis joining the stars and that the stellar wind contribution from the remainder of the hemisphere facing the X-ray source is shut down by X-ray photoionization effects (in both X-ray states). The strong stellar wind from the shadowed side of the supergiant will stall when Coriolis deflection brings the gas into the region of X-ray illumination. This stalled gas component may be overtaken by the orbital motion of the black hole and act to inhibit accretion from the focused wind. The variations in the strength of the shadow wind component may then lead to accretion rate changes that ultimately determine the X-ray state. ", "introduction": "% The massive X-ray binary Cygnus X-1 is a seminal target in the study of gas dynamics in the vicinity of a stellar mass black hole. Its X-ray luminosity and energetic jets \\citep{gal05} are powered by gas accretion from the nearby companion star HD~226868 (O9.7~Iab; \\citealt{wal73}) in a spectroscopic binary with a 5.6 day orbital period. There are several ways in which mass transfer from the supergiant to the black hole may occur in this system \\citep{kap98}. The O-supergiant, like other massive and luminous stars, has a strong radiatively driven wind that may be partially accreted through the gravitational force of the black hole. The supergiant is large and is probably close to filling its critical Roche surface \\citep{gie86a,her95}, so a gas stream through the inner L1 point may also be present. The actual gas flow in the direction of the black hole is probably intermediate between a spherically symmetric wind and a Roche lobe overflow stream, and there is evidence that the flow is best described as a focused wind \\citep{fri82,gie86b,gie03,mil05}. The gas ions responsible for accelerating the wind may become ionized in the presence of a strong X-ray source, leading to a lower velocity, ``stalled'' wind \\citep{blo90,ste91}. In situations of very high X-ray flux, photoionization may extend so close to the supergiant's photosphere that the wind never reaches the stellar escape velocity and thus ceases to become an X-ray accretion source \\citep{day93,blo94}. However, such a high X-ray flux may heat the outer gas layers to temperatures where the thermal velocities exceed the escape velocity to create a thermal wind that may fuel black hole accretion. Important clues about the mass transfer process come from the temporal variations of the observed X-ray flux. Cyg~X-1 is generally observed in either a low flux/hard spectrum state, with an X-ray spectrum that is relatively flat, or a high flux/soft spectrum state with a steeper power-law spectrum \\citep{sha06}. The gamma-ray portion of the spectrum is also elevated during the high/soft state \\citep{mcc02}. The high/soft state usually lasts for periods of days to months, and the fraction of time observed in the high/soft state has increased from $10\\%$ in 1996--2000 to $34\\%$ since early 2000 \\citep{wil06}. This increase may be related to an overall increase in the supergiant's radius in the period from 1997 to 2003 -- 2004 that is suggested by changes in the long term optical light curve \\citep{kar06a}. The system sometimes experiences so-called failed-state transitions, when it starts to increase in flux, but then stops at an intermediate state and returns to the low/hard state. All these transitions probably reflect changes in the inner truncation radius of the accretion disk surrounding the black hole that are caused by a variable accretion rate (largest when the system is in the high/soft state; \\citealt{don02,mcc06}). Thus, the temporal variations in the X-ray state offer us the means to compare the black hole accretion processes with observational signatures related to mass transfer. The H$\\alpha$ emission formed in the high density gas at the base of the stellar wind is an important diagnostic of the mass loss rate in massive stars \\citep{pul96,mar05}. The H$\\alpha$ emission variations in HD~226868 over the last few years are documented in independent spectroscopic investigations by \\citet{gie03} and \\citet*{tar03}. Both of these studies concluded that the H$\\alpha$ emission appears strongest when the system is in the low/hard X-ray state, while a range of weak to moderate emission strengths are observed during the high/soft states. This is a surprising result, since taken at face value, strong emission is associated with a large wind mass loss rate, and the simplest expectation that the X-ray accretion flux increases with mass loss rate is, in fact, not observed. There are several possible explanations: (1) A denser wind may be more opaque to X-rays. However, this seems unlikely because the observed inverse relation between H$\\alpha$ emission strength and X-ray flux is observed at all orbital phases, not just when the supergiant and its wind are in the foreground. (2) The X-ray source may photoionize and heat the gas responsible for the H$\\alpha$ emission, so that a larger X-ray flux leads to a decrease in H$\\alpha$ strength. This clearly occurs at some level, but both \\citet{gie03} and \\citet{tar03} argue that portions of the wind shaded from the ionizing flux also display significant temporal variations. (3) Changes in the X-ray flux will lead to variations in the ionized volume of gas surrounding the black hole, and consequently, the total acceleration of the wind in the direction towards the black hole will vary with the distance traveled before the atoms responsible for line-driving are ionized. Thus, a stronger, denser wind might reach a faster speed before ionization, and since the Bondi-Hoyle accretion rate varies as $\\sim v^{-4}$, the gas captured by the black hole (and the associated X-ray flux) declines. This last process can be tested through direct study of the degree of wind ionization observed in the ultraviolet P~Cygni lines formed in the supersonic part of the wind outflow. When the system is observed with the ionization region in the foreground, the absorption cores of these P~Cygni lines will be truncated at a blueshift corresponding to the highest projected speed before encountering the ionization zone, the so-called Hatchett-McCray effect \\citep{hat77}. The binary is so faint in the ultraviolet that high dispersion spectra were very difficult to obtain with the {\\it International Ultraviolet Explorer (IUE)} satellite \\citep{dav83}, but a good series of observations were obtained with {\\it IUE} at lower spectral resolution that clearly indicate the weakening of the wind lines when the black hole is in front \\citep*{tre80,vlo01}. Most of these spectra were obtained in the low/hard X-ray state (\\S5), when according to the varying wind strength model the mass loss rate is higher and the wind is less ionized and faster. We embarked on a new program of high S/N, high dispersion UV spectroscopy to test this hypothesis with the {\\it Hubble Space Telescope} Space Telescope Imaging Spectrograph (STIS). We obtained observations at the two orbital conjunction phases in both 2002 and 2003. These sets of observations were both made during the rare high/soft X-ray state, and planned observations during the low/hard state were unfortunately scuttled by the STIS electronics failure in 2004. However, we can rebin the high quality {\\it HST} spectra made during the high/soft state to the lower resolution of the {\\it IUE} archival spectra (mostly low/hard state) in order to test whether or not the wind ionization state does in fact differ significantly between states. We describe a program of supporting optical spectroscopy we have obtained to check the orbital phase (\\S2) and wind strength (\\S3) at the times of the {\\it HST} observations. We compare the H$\\alpha$ measurements with the contemporaneous X-ray light curve recorded with the {\\it Rossi X-ray Timing Explorer} All-Sky Monitor instrument \\citep{lev96} and confirm our earlier result showing how H$\\alpha$ tends to strengthen as the X-ray flux declines (\\S3). We then describe the observed variations in the main UV wind lines and present a simple ``shadow wind'' model for the profiles (\\S4). We compare the variations observed in the {\\it HST} and {\\it IUE} spectra in \\S5, and then reassess the question of the mass transfer process in \\S6. We will discuss the photospheric features in the UV spectra in a forthcoming paper (Caballero Nieves et al., in preparation). ", "conclusions": "% The X-ray accretion flux of Cyg X-1 is fueled by mass transfer from the supergiant. We argue in this section that the mass transfer process is dominated by a wind focused along the axis joining the stars. However, the accretion of this gas by the black hole may be influenced by the strength of the radiatively driven shadow wind that is directed away from the black hole. We begin by reviewing the most pertinent observational results from this investigation, and then we consider the interplay between the dynamics of the wind outflow and the X-ray accretion flux. First, the {\\it HST} STIS spectra of HD~226868 that we obtained at two epochs when the system was in the high/soft state show dramatic variations in the wind line strength that result from a superionization of the gas atoms illuminated by the X-ray flux. Shadow wind models, in which the wind ions only exist in the region where X-rays are blocked by the supergiant, make a reasonably good match to the observed profile variations, so we suspect that X-ray photoionization dominates much of the zone between the black hole and the facing hemisphere of the supergiant. A similar degree of wind ionization probably also exists in the X-ray low/hard state since similar orbital variations in wind line strength are found in {\\it IUE} low dispersion spectra made during the X-ray low/hard state. Second, the {\\it HST} spectra suggest that stellar wind gas emanating from parts of the photosphere facing the X-ray source attains only a small velocity before becoming photoionized. For example, the highest optical depth wind feature, \\ion{C}{4} $\\lambda\\lambda 1548, 1550$, shows only a very modest P~Cygni absorption core at phase $\\phi=0.5$ (see Fig.~11) that extends blueward no more than about $-400$ km~s$^{-1}$ (and it is possible that this small component results from a minor part of shadow wind projected against supergiant at $\\phi=0.5$; see Fig.~8). This very low wind speed is probably less than the stellar escape velocity ($\\sim 700$ km~s$^{-1}$ near the poles). These results from the UV wind lines indicate that very little mass loss is occurring by a radiatively driven wind for surface regions that are exposed to the X-ray source. The fact that the wind features appear similarly weak in {\\it IUE} spectra obtained in the low/hard state suggests that a spherical, radiatively driven wind from the hemisphere of the supergiant facing the black hole is probably always weak or absent, and thus, accretion from a spherically symmetric wind must play a minor role in feeding the black hole in Cyg~X-1. On the other hand, we found that the emission equivalent width of the \\ion{He}{2} $\\lambda 4686$ line is consistently strong between X-ray states and over the available record of observation. The orbital phase variations of this spectral feature \\citep{gie86b,nin87} are successfully matched by models of emission from an enhanced density and slower gas outflow region between the stars that is expected for a focused stellar wind \\citep{fri82}. Thus, while X-ray ionization reduces the wind outflow away from the axis joining the stars, the X-ray flux is apparently insufficient to stop the outflow in the denser gas of the focused wind (the result of both tidal and radiative forces). Consequently, it is this focused wind component that is probably the primary means of mass transfer in Cyg~X-1. We caution that the relative constancy of the \\ion{He}{2} $\\lambda 4686$ emission flux does not necessarily imply that the mass loss rate in the focused wind is also steady. For example, the increased X-ray photoionization during the high/soft state may lead to an increase in \\ion{He}{2} $\\lambda 4686$ emission (see Fig.~2 in \\citealt{gie86b}), so that a lower mass loss rate but higher ionization fraction might result in the same amount of observed emission. However, the presence of the \\ion{He}{2} $\\lambda 4686$ emission in both X-ray states indicates that focused wind mass loss always occurs at some level. Finally, we confirm that the H$\\alpha$ P~Cygni line forms mainly in the base of the stellar wind of the supergiant since we observe that H$\\alpha$ follows the orbital velocity curve of the supergiant (Fig.~1). The new observations are consistent with earlier results \\citep{gie03,tar03} in demonstrating that the H$\\alpha$ emission strength is generally weaker in the high/soft X-ray state. Photoionization and heating may extend down to atmospheric levels where the gas densities are sufficient to create H$\\alpha$ emission, so that the reduction in H$\\alpha$ strength in the X-ray high/soft state may partially result from photoionization related processes. However, \\citet{gie03} showed that H$\\alpha$ emission variability was present in those Doppler shifted parts of the profile corresponding to the X-ray shadow hemisphere of the supergiant (see their Fig.~15), so part of the H$\\alpha$ variations must be related to gas density variations at the base of the stellar wind. Thus, the observed H$\\alpha$ variations suggest that the high/soft X-ray state occurs when the global, radiatively driven part of the wind is weaker. Long term, quasi-cyclic variations in wind strength are apparently common among hot supergiants \\citep{mar05}. \\citet{gie03} and \\citet{tar03} suggested that the variations in X-ray state are caused by changes in wind velocity due to changes in supergiant mass loss rate. During times when the supergiant's wind is denser and the mass loss rate is higher, the photoionization region would be more restricted to the region closer to the black hole. Consequently, a radiatively driven wind could accelerate to a higher speed before stalling when the gas enters the ionization zone, and thus, the faster wind would result in a lower black hole accretion rate and X-ray luminosity (creating the low/hard X-ray state). Conversely, if the wind mass loss rate drops, then the X-ray ionization zone will expand, the maximum wind velocity towards the black hole will decline, and the net accretion rate will increase (perhaps creating the high/soft state; \\citealt{ho87}). This creates a positive feedback mechanism that may continue until the wind is ionized all the way down to the stellar photosphere facing the X-ray source \\citep{day93,blo94}. If this scenario is correct, then we expect that the outflow velocities in the direction of the black hole (as measured in blue extent of the P~Cygni lines at phase $\\phi=0.5$) will be larger than the supergiant escape velocity during the X-ray low/hard state. The superb quality {\\it HST}/STIS spectra indicate that outflow velocities are too low to launch the wind during the X-ray high/soft state. Moreover, the low resolution {\\it IUE} spectra from the low/hard state appear to show a very similar pattern of the loss of the P~Cygni absorption at $\\phi=0.5$ (Fig.~12 and 13), indicating that significant ionization zones still exist in the low/hard state. Taken at face value, these {\\it IUE} results suggest that the spherical component of wind outflow towards the black hole is weak and slow in both X-ray states, so a wind speed modulation is probably not the explanation for the accretion variations associated with the X-ray states. We will require new, high quality, UV spectroscopy of Cyg~X-1 during the low/hard state in order to make a definitive test of this idea. The radiatively driven wind of the supergiant leads to effective mass loss only in the X-ray shadowed hemisphere and in the focused wind between the stars in Cyg~X-1. The outflow in the shadow wind region will experience a Coriolis deflection, so that the trailing regions of the shadow wind will eventually enter the zone of X-ray illumination \\citep{blo94}. Once photoionized, this gas will stall with the loss of the important ions for radiative acceleration, and some of this slower gas may extend around the orbital plane to the vicinity of the black hole. Although this deflected wind gas is probably not a major accretion source \\citep{blo94}, it may affect the accretion dynamics of the focused wind. For example, when the shadow wind mass loss rate is high (times of strong H$\\alpha$ emission), the resulting stalled wind component will create a higher ambient gas density on the leading side of the zone surrounding the black hole. The focused wind flow will make a trajectory towards the following side of the black hole, and while gas passing closer to the black hole will merge into an accretion disk, gas further out will tend to move past the black hole before turning into the outskirts of the disk. The presence of the stalled gas on the leading side may deflect away this outer, lower density part of the flow and effectively inhibit gas accretion from the focused wind. The subsequent reduction in gas accretion by the black hole may correspond to the conditions required to produce the low/hard X-ray state, while conversely a reduction in the stalled gas from the shadow wind may promote mass accretion and produce the high/soft state \\citep{bro99,don02,mcc06}. Clearly, new hydrodynamical simulations are needed to test whether the stalled wind component is sufficient to alter the accretion of gas from the focused wind and create the environments needed for the X-ray transitions." }, "0801/0801.4415_arXiv.txt": { "abstract": "{ We show that the entropy created by Ohmic dissipation inside an accreting charged black hole may exceed the Bekenstein-Hawking entropy by a large factor. If the black hole subsequently evaporates, radiating only the Bekenstein-Hawking entropy, then the black hole appears to destroy entropy, violating the second law of thermodynamics. A companion paper discusses the implications of this startling result. Bousso's covariant entropy bound is not violated. } ", "introduction": "The purpose of this paper is to show that classical processes of dissipation can generate huge quantities of entropy inside the horizon of a black hole, many orders of magnitude more than the Bekenstein-Hawking \\cite{Bekenstein:1973ur} entropy. The specific black hole model presented is intended to be semi-realistic, albeit over-simplified, with parameters appropriate to a real supermassive black hole. We take charge as a surrogate for angular momentum, and electrical conductivity as a surrogate for angular momentum transport. To see how much entropy might be created, we treat the electrical conductivity as an adjustable free parameter. If a black hole creates many times the Bekenstein-Hawking entropy and subsequently evaporates, radiating only the Bekenstein-Hawking entropy and leaving no remnant, then entropy is destroyed, violating the second law of thermodynamics. The implications of this startling result are discussed in a companion paper \\cite{Polhemus:2009}. Throughout this paper we treat entropy in a purely classical fashion. In particular, we assume that locality holds inside the black hole. Locality, the quantum field theory proposition that operators commute at spacelike-separated points, is the assumption that normally makes it legitimate to add entropy over spacelike surfaces. Since the spacetime curvature inside a supermassive black hole is well below Planck, except near the singularity, one might expect classical physics to apply. It is widely thought that in order to preserve unitarity of black hole evaporation, locality must break down over spacelike surfaces connecting the inside and outside of a black hole \\cite{Susskind:1993if}. In the companion paper \\cite{Polhemus:2009} we argue that the calculations of the present paper point to a profligate breakdown of locality inside black holes. We work in Planck units, $k_B = c = G = \\hbar = 1$. ", "conclusions": "\\label{summary} We have shown that the dissipation of the free energy of the electric field inside a charged black hole can potentially create many times more than the Bekenstein-Hawking entropy. If the black hole subsequently evaporates, radiating only the Bekenstein-Hawking entropy and leaving no remnant, then entropy is destroyed. This startling conclusion is premised on the assumption that entropy created inside a black hole accumulates additively on spacelike slices, which in turn derives from the assumption that the Hilbert space of states is multiplicative over spacelike-separated regions, as postulated by locality. This is essentially the same reasoning that originally led Hawking \\cite{Hawking:1976ra} to conclude that black hole evaporation is non-unitary, that black holes must destroy information. It is widely thought that unitarity should be considered a higher principle than locality. To ensure that black hole evaporation is unitary, locality between the inside and outside of a black hole must break down. In a companion paper \\cite{Polhemus:2009} we argue that the gross violation of the second law found in the present paper points to a wholesale breakdown of locality inside black holes, and provides a compelling argument in favor of the conjecture of ``observer complementarity''. The black hole respects Bousso's covariant entropy bound \\cite{Bousso:2002ju}, as it should given the theorem of \\cite{Flanagan:1999jp}." }, "0801/0801.4309_arXiv.txt": { "abstract": "We present a new method for the reconstruction of the longitudinal profile of extensive air showers induced by ultra-high energy cosmic rays. In contrast to the typically considered shower size profile, this method employs directly the ionization energy deposit of the shower particles in the atmosphere. Due to universality of the energy spectra of electrons and positrons, both fluorescence and Cherenkov light can be used simultaneously as signal to infer the shower profile from the detected light. The method is based on an analytic least-square solution for the estimation of the shower profile from the observed light signal. Furthermore, the extrapolation of the observed part of the profile with a Gaisser-Hillas function is discussed and the total statistical uncertainty of shower parameters like total energy and shower maximum is calculated. ", "introduction": "The particles of an extensive air shower excite nitrogen molecules in the atmosphere, which subsequently radiate ultraviolet fluorescence light isotropically. This fluorescence light signal can be measured with appropriate optical detectors such as the fluorescence telescopes of HiRes~\\cite{fd:HiRes}, the Pierre Auger Observatory~\\cite{fd:Auger} or the Telescope Array~\\cite{fd:TA}. The number of emitted fluorescence photons is expected to be proportional to the energy deposited by the shower particles. Recent measurements of the fluorescence yield in the laboratory confirm this expectation within the experimental uncertainties~\\cite{Kakimoto:1995pr,Waldenmaier:2007am,airflydEdX}. Non-radiative processes of nitrogen molecule de-excitation lead to a temperature, pressure and humidity dependence of the fluorescence yield (see e.g.~\\cite{airflyAtm}). For atmospheric parameters of relevance to the reconstruction of air showers of ultra-high energy cosmic rays, the pressure dependence of the ionization energy deposit per meter track length of a charged particle is almost perfectly canceled by the pressure dependence of the fluorescence yield (see, for example, \\cite{Keilhauer:2005nk}). Therefore only a weak pressure and temperature dependence has to be taken into account if the number of emitted photons is converted to a number of charged particle times track length, as has been done in the pioneering Fly's Eye experiment \\cite{Baltrusaitis:1985mx}. The reconstructed longitudinal shower profile is then given by the number of charged particles as function of atmospheric depth. The approximation of assuming a certain number of fluorescence photons per meter of charged particle track and the corresponding expression of the longitudinal shower development in terms of shower size are characterized by a number of conceptual shortcomings. Firstly the energy spectrum of particles in an air shower changes in the course of its development. A different rate of fluorescence photons per charged particle has to be assumed for early and late stages of shower development as the ionization energy deposit depends on the particle energy \\cite{Song:1999wq}. Secondly the tracks of low-energy particles are not parallel to the shower axis leading to another correction that has to be applied \\cite{Alvarez-Muniz:2003my}. Thirdly the quantity ``shower size'' is not suited to a precise comparison of measurements with theoretical predictions. In air shower simulations, shower size is defined as the number of charged particles above a given energy threshold $E_{\\rm cut}$ that cross a plane perpendicular to the shower axis. Setting this threshold very low to calculate the shower size with an accuracy of $\\sim 1$\\% leads to very large simulation times as the number of photons diverges for $E_{\\rm cut}\\rightarrow 0$. Moreover, the shower size reconstructed from data depends on simulations itself since the shower size is not directly related to the fluorescence light signal. These conceptual problems can be avoided by directly using energy deposit as the primary quantity for shower profile reconstruction as well as comparing experimental data with theoretical predictions. Due to the proportionality of the number of fluorescence photons to the energy deposit, shower simulations are not needed to reconstruct the total energy deposit at a given depth in the atmosphere. Another advantage is that the calorimetric energy of the shower is directly given by the integral of the energy deposit profile \\cite{linsleydEdX}. Furthermore the energy deposit profile is a well-defined quantity that can be calculated straight-forwardly in Monte Carlo simulations and does not depend on the simulation threshold \\cite{Risse:2003fw}. Most of the charged shower particles travel faster than the speed of light in air, leading to the emission of Cherenkov light. Thus, in general, the optical signal of an air shower consists of both fluorescence and Cherenkov light contributions. In the traditional method \\cite{Baltrusaitis:1985mx} for the reconstruction of the longitudinal shower development, the Cherenkov light is iteratively subtracted from the measured total light. The drawbacks of this method are the lack of convergence for events with a large amount of Cherenkov light and the difficulty of propagating the uncertainty of the subtracted signal to the reconstructed shower profile. An alternative procedure, used in \\cite{Abbasi:2004nz}, is to assume a functional form for the longitudinal development of the shower, calculate the corresponding light emission and vary the parameters of the shower curve until a satisfactory agreement with the observed light at the detector is obtained. Whereas in this scheme the convergence problems of the aforementioned method are avoided, its major disadvantage is that it can only be used if the showers indeed follow the functional form assumed in the minimization. It has been noted in~\\cite{cher:giller} that, due to the universality of the energy spectra of the secondary electrons and positrons within an air shower, there exists a non-iterative solution for the reconstruction of a longitudinal shower profile from light detected by fluorescence telescopes. Here we will present an analytic least-square solution for the estimation of the longitudinal energy deposit profile of air showers from the observed light signal, in which both fluorescence and Cherenkov light contributions are treated as signal. We will also discuss the calculation of the statistical uncertainty of the shower profile, including bin-to-bin correlations. Finally we will introduce a constrained fit to the detected shower profile for extrapolating it to the regions outside the field of view of the fluorescence telescope. This constrained fit allows us to always use the full set of profile function parameters independent of the quality of the detected shower profile. \\begin{figure*} \\begin{center} \\includegraphics[clip,width=0.95\\textwidth]{sketch.eps} \\caption{Illustration of the isotropic fluorescence light emission (solid circles), Cherenkov beam along the shower axis (dashed arcs) and the direct (dashed lines) and scattered (dotted lines) Cherenkov light contributions.}\\label{fig1} \\end{center} \\end{figure*} ", "conclusions": "In this paper a new method for the reconstruction of longitudinal air shower profiles was presented. With the help of simulations we have shown that the least square solution yields robust and unbiased results and that uncertainties of shower parameters can be reliably calculated for each event.\\\\ Events with a large Cherenkov light contribution are currently usually rejected during the data analysis (see for instance \\cite{Abbasi:2004nz,AbuZayyad:2000ay}) However, as we have shown, there is no justification for rejecting such showers, once experimental systematic uncertainties are well understood. Because events with a large Cherenkov contribution have different systematic uncertainties to those dominated by fluorescence light, both event classes can be compared to study their compatibility.\\\\ At energies below 10$^{17.5}$ \\eV{}, where new projects \\cite{heat,tale} are planned to study the transition from galactic to extragalactic cosmic rays, events with a large fraction of direct Cherenkov light will dominate the data samples, because the amount of light, and thus trigger probability, of these events is much larger than that of a fluorescence dominated shower. If at these energies it is still possible to measure an accurate shower geometry, the fluorescence detectors should in fact be used as Cherenkov-Fluorescence telescopes.\\\\" }, "0801/0801.4637_arXiv.txt": { "abstract": "This paper addresses the origin of the silicate emission observed in PG QSOs, based on observations with the {\\it Spitzer Space Telescope}. Scenarios based on the unified model suggest that silicate emission in AGN arises mainly from the illuminated faces of the clouds in the torus at temperatures near sublimation. However, detections of silicate emission in Type 2 QSOs, and the estimated cool dust temperatures, argue for a more extended emission region. To investigate this issue we present the mid-infrared spectra of 23 QSOs. These spectra, and especially the silicate emission features at $\\sim 10$ and $\\sim 18$ $\\mu$m, can be fitted using dusty narrow line region (NLR) models and a combination of black bodies. The bolometric luminosities of the QSOs allow us to derive the radial distances and covering factors for the silicate-emitting dust. The inferred radii are 100-200 times larger than the dust sublimation radius, much larger than the expected dimensions of the inner torus. Our QSO mid-IR spectra are consistent with the bulk of the silicate dust emission arising from the dust in the innermost parts of the NLR. ", "introduction": "Unified schemes for active galactic nuclei (AGN) postulate an obscuring torus surrounding an accreting super-massive black hole. Models predict that the infrared spectral energy distribution (SED) of the torus depends sensitively on its orientation, geometry and density distribution \\citep[e.g.][]{pier92,granato94,efstathiou95,granato97,nenkova}. In particular, the tori are predicted to exhibit prominent silicate dust features in either absorption or emission, depending on whether an AGN is viewed with the torus edge-on (Type 2) or face-on (Type 1). Previous failures to detect strong 9.7$\\mu$m silicate emission in Type 1 AGN led to several proposed modifications of the unified model. For example modified grain size distributions have been assumed \\citep{laordrain1993,maiolino} or a clumpiness of the torus invoked \\citep{nenkova}. The {\\it Spitzer Space Telescope} ({\\it Spitzer}), with its good mid-infrared (mid-IR) wavelength coverage and sensitivity, has drastically changed our view of this problem. \\citet{siebenmorgen05a} and \\citet{hao05} reported the first \\spitzer\\ Infrared Spectrograph (IRS) detections of prominent silicate emission features in the mid-infrared spectra of several luminous quasars. Sturm et al. (2005) reported the first detection of 10 and 18 $\\mu$m silicate emission features in a low-luminosity LINER (NGC 3998). Comparison to the 10/18 $\\mu$m feature ratio of optically thin emission from silicate dust at different temperatures suggests a modest temperature ($\\sim$200K) of the emitting dust. The presence of prominent 10$\\mu$m silicate emission features in AGN covering a broad range in luminosity could be taken as direct evidence for the existence of an obscuring torus. However, it is not at all clear whether this emission actually arises from the inner regions of a face-on torus. Depending on the size and composition of the grains, sublimation occurs between $\\sim$~800 and 1500 K \\citep{kimura} which is also the temperature range expected for silicate dust located near the hot inner torus wall. The lower temperature indicated for the emitting silicate dust can be interpreted as evidence for dust emission from regions located further away from the central heating source. Several arguments support such a scenario. Silicate emission has also been detected in Type-2 QSOs \\citep{sturm2006,teplitz06}, whereas for an edge-on view of the torus, one would expect to see silicate in absorption only. An extended emitting region, with dimensions much larger than the inner torus dimension, is fully consistent with this result. Broad-band $10\\mu m$ imaging of several nearby AGN suggests extended mid-infrared continuum \\citep{cameron93,tomono,bock00,radomski03,packham05}. \\citet{efstathiou06} has modeled the silicate emission of the Type-2 QSO IRASF10214+4724 \\citep{teplitz06} invoking extended NLR dust in addition to an AGN torus. \\citet{marshall07} conclude that some of the optically thin warm emission in the QSO PG0804+761 may emerge from regions beyond the torus and suggest clouds in the NLR as a possible origin of this emission. These arguments indicate that silicate emission may originate in extended regions ($\\sim 100$ pc dimension). The nature and location of the extended silicate-emitting region is not yet known. In this paper we explore in a quantitative way one plausible interpretation, namely the association of this cool dust with the NLR. To this end, we present fits of our QSO spectra with a superposition of NLR dust models and spectral components representing the innermost hot dust and the bulk of the inner structure emission (both related to the torus) as well as the large scale host emission. The fitted model and the bolometric luminosity of each source enable us to estimate the cool dust distance and its covering factor. We note that our models cannot exclude the possibility of a torus contribution to the observed silicate emission. In section \\ref{sec:Obs} we describe our QSO sample. In section \\ref{sec:model}, we introduce the model components and detail our fitting procedure. We also describe how we estimate the silicate dust cloud distances and the related covering factors. In section \\ref{sec:robustness}, we discuss the dependence on model parameters. In section \\ref{sec:results}, we present the results of the fits, which are then discussed in section \\ref{sec:discussion}. Finally section \\ref{sec:conclusions} summarizes our conclusions. ", "conclusions": "\\label{sec:conclusions} Scenarios based on the unified model suggest that silicate emission in AGN arises mainly from the illuminated faces of the clouds in the torus at temperatures near sublimation. However, detections of silicate emission in Type 2 QSOs, and the estimated cool dust temperatures, argue for an origin in a more extended region. To investigate this issue, we have presented the \\spitzer-IRS spectra of 23 QSOs. We have matched physically-based models to the mid-infrared spectra and found that the silicate emission observed in these objects can be reproduced by emission from clouds, outside the central torus. This extended silicate-emitting region is possibly associated with the innermost NLR region or the intermediate dusty region proposed by \\citet{Netzer93}. The dust cloud distances found here scale with the AGN luminosity as $R_{dust}\\sim 80\\cdot L_{bol46}^{0.5}$ pc, with the bolometric luminosity $L_{bol46} $ given in units of $10^{46}$ $erg s^{-1}$. We have estimated the median distance of the dust cloud responsible for the silicate emission to be 40 pc, while for individual sources distances up to 260 pc are possible. The smallest cloud distance is 9 pc for PG 0050+124. The calculated covering factors for the dust clouds have a median of 0.16 and are in agreement with an NLR origin of the silicate emission. Our models do not exclude the possibility of a torus contribution to the observed silicate emission, but rather emphasize the good agreement and perhaps the necessity of a larger-scale contribution to this emission. Finally, future high spatial resolution infrared observations and further crosschecks including the comparison to Type 1/2 ratios as well as distributions and individual values for the obscuring columns in X-rays and the near infrared are needed to resolve the remaining ambiguities between torus and NLR emission." }, "0801/0801.3684_arXiv.txt": { "abstract": "Kant and Laplace suggested the Solar System formed from a rotating gaseous disk in the 18th century, but convincing evidence that young stars are indeed surrounded by such disks was not presented for another 200 years. As we move into the 21st century the emphasis is now on disk formation, the role of disks in star formation, and on how planets form in those disks. Radio wavelengths play a key role in these studies, currently providing some of the highest spatial resolution images of disks, along with evidence of the growth of dust grains into planetesimals. The future capabilities of EVLA and ALMA provide extremely exciting prospects for resolving disk structure and kinematics, studying disk chemistry, directly detecting proto-planets, and imaging disks in formation. ", "introduction": "The search for forming planetary systems is ultimately a search for our origins. How do the gas and dust in molecular clouds evolve into rocky and gas giant planets, and how common are planetary systems like the Solar System? Radial velocity searches for extra solar planets indicate that the Solar System might actually be unusual, but such searches are only just achieving the sensitivities needed to detect the Solar System-like planetary systems. For several reasons it is much easier to study the disks from which planetary systems are expected to form, and this may shed light on the environment and processes taking place in the early Solar System. The formation of disks at some stage during the star formation process is made inevitable by one of several ``angular momentum problems'' in star formation. Specific angular momentum is imparted to molecular clouds by differential Galactic rotation, and many orders of magnitude must be lost before a planetary system can form. The outcome of the star formation process for solar-type stars is therefore a T Tauri star, descending the Hayashi track towards the main sequence in the HR diagram, surrounded by a disk. Massive stars reach the main sequence while still deeply embedded and accreting. A combination of radiative heating by the central star, accretion, and heating by the interstellar radiation field results in temperatures for disks on Solar System size scales of 10 to 20~K for Solar-type stars. The bulk of the disk mass is therefore best traced by millimeter and submillimeter wavelength emission. Indeed, the first convincing evidence for the ubiquity of disks surrounding T Tauri stars arose from a single-dish survey for 1.3~mm continuum emission by \\citet{chandler:Beckwith1990}. Although those measurements were not able to resolve the emission the flux densities detected, and corresponding dust column densities, implied that the dust had to be distributed in flattened structures in order to explain the low optical extinctions to the central stars. The rest of this review focuses on the role of radio interferometry in spatially resolving the emission from circumstellar disks, and in understanding the disk properties. ", "conclusions": "Over the last few years, submm, mm, and cm-wave observations of circumstellar disks around low-mass, pre-main sequence stars have demonstrated evidence for grain growth to cm-sized particles, a necessary precursor to planet formation. Keplerian velocity fields in those disks have enabled independent measurements of central stellar masses, and are able to constrain pre-main sequence evolutionary tracks. Deviations from Keplerian rotation have been observed in some cases, and raise the possibility that low-mass companions may be inferred from such observations. Evidence for angular momentum transport in disk winds has been demonstrated by the first observations of rotation in stellar jets, in a sense consistent with that of the accompanying disk. Masers are now able to trace disk and wind interactions for a massive protostar, suggesting that disk-mediated accretion may be a mechanism for forming high-mass stars. The future of disk studies is very bright; ALMA and EVLA will transform this field by providing images of gaps in disks, detecting dust heated by proto-planets, revealing the early phases of disk formation, and through detailed studies of the gas chemistry." }, "0801/0801.3367_arXiv.txt": { "abstract": "In this paper we give a review of the Bowen fluorescence survey, showing that narrow emission lines (mainly N\\,III and C\\,III lines between 4630 and 4660 \\AA) appear to be universally present in the Bowen blend of optically bright low mass X-ray binaries. These narrow lines are attributed to reprocessing in the companion star giving the first estimates of $K_2$, and thereby providing the first constraints on their system parameters. We will give an overview of the constraints on the masses of the compact objects and briefly highlight the most important results of the survey. Furthermore, we will point out the most promising systems for future follow-up studies and indicate how we think their estimates of the component masses can be improved. ", "introduction": "One of the main aims of optical observations of low-mass X-ray binaries (LMXBs) has always been to find a signature of the donor star, and thereby constrain the masses of both components. Unfortunately, the optical emission of most LMXBs that accrete at high rates is completely dominated by the reprocessing of the X-rays in the outer accretion disk. This is the reason why, in spite of optical counterparts being known for persistent LMXBs for years (sometimes even over 20 years), no strong constraints on their system parameters existed until recently. \\begin{figure*}[t] \\parbox{8.0cm}{\\psfig{figure=sco.ps,width=8.0cm}} \\parbox{8.0cm}{\\psfig{figure=doppler.eps,angle=-90,width=8.0cm}} \\caption{{\\it Left:} Trailed spectrogram of the Bowen blend and He\\,II $\\lambda$4686 of Sco\\,X-1 showing the presence and the movement of the narrow emission lines. From \\citet{sc02}. {\\it Right:} Example Doppler maps of He\\,II $\\lambda$4686 (left) and the Bowen region (right) for 4U\\,1636$-$536 (top) and 4U\\,1735$-$44 (bottom). The He\\,II maps trace the accretion disk, while the Bowen maps are dominated by a compact spot at the phasing and velocity where the companion star is expected. The Roche lobe of the secondary and the gas stream trajectory are overplotted for clarity. From \\citet{ccs06}. \\label{scox1}} \\end{figure*} The discovery of narrow high-excitation emission lines in the bright LMXB Sco\\,X-1 opened new opportunities for system parameter constraints in active X-ray binaries \\citep{sc02}. Phase-resolved blue spectroscopy showed that these narrow lines moved in anti-phase with the compact object, strongly suggesting that they arise from the irradiated surface of the donor star. This lead to the first estimate of the semi-amplitude of the radial velocity of the secondary, $K_2$, and mass function ($f$($M_1$)=$K_2^3$$P_{\\rm orb}$/4$\\pi$$G$) of Sco\\,X-1. These lines were most prominent in the Bowen region (the region from 4630 to 4660 \\AA), that mainly consists of a blend of N\\,III and C\\,III lines. The N\\,III lines are the result of a UV fluorescence process, while the C\\,III lines are due to photo-ionization and subsequent recombination \\citep{mct75}. In this paper we review the results of a survey we performed to apply this new technique to other persistent LMXBs or transients during outburst in order to find a donor star signature. ", "conclusions": "We have given a short overview of the novel technique of Bowen fluorescence that has proven itself to work for the 9 systems thus far observed (see Table\\,\\ref{limits}), and for most of the systems we have been able to derive the first constraints on their system parameters ever. There are still a handful of persistent LMXBS left that are bright enough for this technique, and there are plans to also observe these systems. Recently 4U\\,1957$+$11 has been observed with Magellan, and in the near future EXO\\,0748$-$676, Ser\\,X-1 and 4U\\,1556$-$605 will be observed with the VLT. Furthermore, this technique might also be an excellent tool to determine the system parameters of future bright transients. The next step will be to improve the determined $K_2$ velocities for the most promising candidates by measuring the distribution of the emission across the Roche-lobe using high resolution spectroscopy. This will allow us to more accurately measure the radial velocities of the narrow lines, determine their rotational broadening, and hopefully model the shape of the lines as a function of orbital period in order to get a better constraint on the $K$-correction. Furthermore, we have started using the fact that the Bowen lines are efficiently reprocessed on the donor star to constrain the inclination and the $K$-correction using Echo-tomography with a narrow band filter centered around the Bowen blend (for more information see Mu\\~noz-Darias et~al. in this volume). These next steps will hopefully provide stronger constraints on the masses of LMXBs, and in particular give accurate neutron star masses." }, "0801/0801.1681_arXiv.txt": { "abstract": "We report CCD photometry of the dwarf nova BF Ara throughout fifteen consecutive nights in quiescence. Light curve in this interval is dominated by a large amplitude ($\\sim 0.8$ mag) modulation consisting two periods. Higher amplitude signal is characterized by period of 0.082159(4) days, which was increasing at the rate of $\\dot P/P_{\\rm sh} = 3.8(3)\\cdot 10^{-5}$. Weaker and stable signal has period of 0.084176(21) days. Knowing the superhump period of BF Ara determined by Kato et al. (2003) and equal to 0.08797(1) days, the first modulation is interpreted as quiescent negative superhump arising from retrograde precesion of titled accretion disk and the latter one as an orbital period of the binary. The respective period excess and defect are $\\epsilon_+ = 4.51\\% \\pm 0.03\\%$ and $\\epsilon_- = -2.44\\% \\pm 0.02\\%$. Thus BF Ara is yet another in-the-gap nova with mass ratio of around $q\\approx0.21$. \\noindent {\\bf Key words:} Stars: individual: BF Ara -- binaries: close -- novae, cataclysmic variables ", "introduction": "The SU UMa variables are a subclass of dwarf novae characterized by the presence of two types of eruptions - normal outburst having an amplitude of 2-3 mag and repeating typically every few tens of days and superoutburts with amplitude of 3-4 mag, lasting few times longer and occuring every year or so. These stars are belived to be binary systems consiting of the white dwarf primary and low mass main-sequence secondary filling its Roche lobe and loosing material which forms an accretion disc around the primary. Almost all SU UMa stars are close binaries with orbital period from 1.25 to slightly over 2 hours. This behaviour in now quite well understood within the frame of the thermal-tidal instability model (see Osaki 1996 for review). Normal outbursts are caused by the thermal instability connected with transition of the material in the disc from neutral to ionized state. Superoutbursts are a result of combined thermal and tidal instability, which, working together, are very effective with sweeping out the matter from the disc. During the superoutburst, the characteristic tooth-shape light modulations with a period a few percent longer than the orbital period of the binary are observed. They are most probably the result of accretion disc \"precession\" (in fact it is not classical precession but change of the position of the line of apsides) caused by gravitational perturbations from the secondary. These perturbations are most effective when disc particles moving in eccentric orbits enter the 3:1 resonance. Then the superhump period is simply the beat period between orbital and \"precession\" rate periods. At the beginning of the 1990s, the quite simple class of SU UMas started to be more complicated. The systems showing superhumps were divided into four subgroups: \\begin{itemize} \\item WZ Sge stars, characterized by an extremely long quiescent state, going into superoutburst every $\\sim$10 years and showing no or very infrequent ordinary outbursts (Patterson et al. 2002), \\item ordinary SU UMa stars, \\item ER UMa stars - systems characterized by an extremely short supercycle (20-60 days), a short interval between normal outbursts (3-4 days) and small amplitude (2--3 mag) of superoutbursts (Kato and Kunjaya 1995, Robertson et al. 1995), \\item permanent superhumpers - high accretion rate systems being permanently in superoutbursts (Skillman and Patterson 1993). \\end{itemize} \\smallskip We now believe that the classes listed above represent increasing mass transfer. Systems with mass transfer rates as low as $10^{15}$ g/s are inactive, quiet and evolved stars of WZ Sge type containing brown dwarf degenerated secondary. Classical SU UMa stars have mass transfer rates one magnitude higer and go into the superoutburt every year or so. ER UMa stars are high mass transfer rate systems (few times per $10^{16}$ g/s) showing frequent and relatively long superoutburst lasting for about half of the supercycle. The explanation of the shortest supercycles involves poorly understood mechanisms causing premature quench of the eruption (Osaki 1995). And finally, permanent superhumpers are system with accretion disc which is thermaly stable and tidaly unstable all the time, i.e. being in permanent state of superoutburst. ", "conclusions": "Our main conclusions may be summarized as follows: \\begin{itemize} \\item Light curve of BF Ara in quiescence is dominated by high amplitude signal with mean period of 0.082159(4) days which during two weeks observing run increased at the rate of $\\dot P/P_{\\rm sh} = 3.8(3)\\cdot 10^{-5}$ \\item We interpret this periodicity as being due to a negative superhump. Such an interpretation is confirmed by the location of BF Ara in the Ritter et al. (2002a) relation, \\item BF Ara seems to be a twin of V503 Cyg (Harvey et al., 1995) - both stars are very active (supercycles of 80--90 days), both show large amplitude negative superhumps in quiescence characterized by changing value period, \\item Prewhitening of the original light curve with the main periodicity resulted in the discovery of another modulation with constant period equal to 0.084176(21) days, which is interpreted as the orbital period of the system. This value makes BF Ara another in-the-gap cataclysmic variable, \\item Knowing the ordinary superhump period measured by Kato et al. (2003) we were able to calculate the period excess as equal to $4.51\\% \\pm 0.03\\%$ which indicates the mass ratio of $q\\approx0.21$. \\end{itemize} \\bigskip \\noindent {\\bf Acknowledgments.} ~We acknowledge generous allocation of the SAAO 1-m telescope time. We would like to thank Prof. J\\'ozef Smak for fruitful discussions. This work was supported for SALT grant number 76/E-60/SPB/MSN/P-03/DWM 35/2005-2007." }, "0801/0801.3198_arXiv.txt": { "abstract": "The energy spectra of TeV gamma-rays from blazars, after being corrected for intergalatic absorption in the Extragalactic Background Light (EBL), appear unusually hard, a fact that poses challenges to the conventional models of particle acceleration in TeV blazars and/or to the EBL models. In this paper we show that the internal absorption of gamma-rays caused by interactions with dense narrow-band radiation fields in the vicinity of compact gamma-ray production regions can lead to the formation of gamma-ray spectra of an almost arbitrary hardness. This allows significant relaxation of the current tight constraints on particle acceleration and radiation models, although at the expense of enhanced requirements to the available nonthermal energy budget. The latter, however, is not a critical issue, as long as it can be largely compensated by the Doppler boosting, assuming very large ($\\geq 30$) Doppler factors of the relativistically moving gamma-ray production regions. The suggested scenario of formation of hard gamma-ray spectra predicts detectable synchrotron radiation of secondary electron-positron pairs which might require a revision of the current ``standard paradigm'' of spectral energy distributions of gamma-ray blazars. If the primary gamma-rays are of hadronic origin related to $pp$ or $p \\gamma$ interactions, the ``internal gamma-ray absorption'' model predicts neutrino fluxes close to the detection threshold of the next generation high energy neutrino detectors. ", "introduction": "The recent reports on detections of very high energy (VHE) gamma-rays from blazars with redshifts $z \\geq 0.1$ (for a review see e.g. Hinton (2007)) initiated renewed debates on the interpretation of TeV gamma-ray spectra of blazars, in particular in the context of the level of the diffuse extragalactic background radiation at optical and infrared wavelengths, often called also as Extragalactic Background Light (EBL). Initially, the tight link between these two topics - TeV blazars and EBL - became a subject of hot discussions prompted by multi (up to 20) TeV gamma-rays detected from a nearby BL Lac object, Mkn~501 \\cite{mkn501_HEGRA}, and by the reports claiming detection of high fluxes of EBL at far infrared wavelengths \\cite{Hauser,Schlegel,Lagage,Fink}. However, it was quickly recognised that these two claims hardly could be compatible within any standard model of TeV blazars (see, for a review, Aharonian (2001)). A distinct feature of extragalactic gamma-ray astronomy is that VHE gamma-rays emitted by distant ($\\geq 100 \\ \\rm Mpc$) objects arrive after significant absorption caused by their interactions with EBL via the process $\\gamma \\gamma \\rightarrow e^+ e^-$ \\cite{Nikishov,Jelley,Gould}. The reconstructed, i.e. the absorption-corrected gamma-ray spectrum from a source at a redshift z, $J_0(E)=J_{\\rm obs}(E) e^{\\tau(E,z)}$ % depends on the flux and energy spectrum of EBL through the optical depth $\\tau(E,z)$. Thus, at energies where $\\tau(E,z) \\geq 1$, the primary gamma-rays suffer strong spectral deformation. The EBL consists of two emission components produced by stars and partly absorbed/re-emitted by dust throughout the entire history of galaxy evolution. As a result, two distinct bumps are expected in the spectral energy distribution (SED) of EBL at near infrared (NIR) and far infrared (FIR) wavelengths, with a mid-infrared (MIR) ``valley'' between these two bumps (see e.g. Hauser and Dwek (2001)). Generally, for almost all EBL models, $\\tau(E)$ is a strong function of energy below 1 TeV and above 10 TeV; between 1 and 10 TeV the energy-dependence of $\\tau(E)$ is much weaker \\cite{Hamburg01}. Consequently, one should expect significant distortion of the VHE spectra of blazars at energies below 1 TeV and above 10 TeV, provided that at these energies $\\tau \\geq 1$. One can re-formulate this statement in a different way. Namely, for a standard (``decent'') intrinsic gamma-ray spectrum, the observer should detect very soft (steep) spectra at energies below 1 TeV and above 10 TeV from objects for which $\\tau \\geq 1$ at corresponding energies. This condition is safely satisfied, given the constraints on the minimum EBL flux imposed by galaxy counts, for blazars with redshifts $z \\geq 0.15$ like 1ES~1101-232 and for nearby objects with $z \\sim 0.03$ like Mkn~501. Even though the \\textit{detected} gamma-ray spectra from both objects in the corresponding energy intervals are indeed quite steep with a photon index $\\sim 3$ \\cite{mkn501_HEGRA,1101_HESS}, they appear not sufficiently steep to compensate the function $f(E)=e^{\\tau(E)}$, and thus to prevent a robust conclusion that the \\textit{intrinsic} VHE gamma-ray spectra of these blazars are unusually hard. In the case of Mkn 501, the intrinsic spectrum has a ``non-standard'' shape with a possible pile-up above 10 TeV which has been interpreted as a ``IR background - TeV gamma-ray crisis'' \\cite{IRcrisis} or a need to invoke dramatic assumptions like a violation of the Lorentz invariance (see e.g. Kifune (1999)). However, a more pragmatic view which presently dominates in both infrared and gamma-ray astronomical communities, treats this ``crisis'' as somewhat exaggerated, especially given the ambiguity of extraction of the truly diffuse extragalactic FIR component from the much higher backgrounds of local origin (see e.g. Hauser and Dwek (2001)). Nevertheless, the recently reported low limits on the EBL at mid infrared wavelengths from the Spitzer deep cosmological surveys appeared quite high, for example at 70 $\\mu \\rm m$ the EBL flux should exceed $\\geq 7.1 \\pm 1.0 \\ \\rm nW/m^2 s$ \\cite{Dole}. This implies that the problem is not yet over, and one may still face a challenge with the interpretation of the energy spectra of Mkn 501 and Mkn 421 in the multi-TeV energy domain. On the other hand, the recent detections of TeV gamma-rays from blazars with redshifts $z \\geq 0.15$ renewed the potential problems and challenges for standard models of TeV blazars. This time the issue has a more solid experimental background, because the gamma-ray spectra corrected for the intergalactic absorption appear very hard (``harder than should be'') even for the minimum possible EBL fluxes at optical and NIR wavelengths. Namely, the HESS collaboration reported, based on the detection of TeV gamma-rays from the BL Lac object 1ES~1101-232, that any significant deviation from the lower limits of EBL determined by the integrated light of galaxies resolved by the Hubble telescope \\cite{Madau}, would lead to very hard intrinsic gamma-ray spectrum with a slope characterized by a photon index $\\Gamma_0 \\leq 1.5$ \\cite{1101_HESS}. The analysis based on a larger sample of TeV blazars leads to the same conclusion \\cite{Mazin}. Recently, the HESS collaboration reported detection of multi-TeV gamma-rays from 1ES 0229+200, a BL Lac object located at a redshift z=0.1396 \\cite{0229_HESS}. It is remarkable that the \\textit{detected} hard gamma-ray spectrum of this source with a photon index $\\Gamma_{\\rm obs} \\sim 2.5$ extends up to 15 TeV. This, to a certain extent surprising result can be explained by the shape of the energy flux of EBL which between the NIR and MIR bands is expected to be proportional to $\\lambda^{-1}$ \\cite{Hamburg01}. Yet, the absolute EBL flux, derived from a rather conservative assumption that the photon index of the intrinsic spectrum of TeV gamma-rays does not exceed $1.5$, appears again close to the EBL lower limit, this time at MIR ($\\approx 2-3 \\ \\rm nW/m^2 sr$ at $10 \\ \\rm \\mu m$), derived from the Spitzer galaxy counts \\cite{Fazio,Dole}. Thus, the gamma-ray observations of 1ES~1101-232 and 1ES 0229+200 can be interpreted as an argument that the galaxies resolved by the Hubble and Spitzer telescopes provide the bulk of the EBL flux from optical to mid infrared wavelengths. Given the importance of such a statement, in particular for understanding of contribution of the first stars to the EBL (see e.g. Kashlinsky (2005); Mapelli et al. (2006)), it is essential to explore alternative ways of explanation of very hard intrinsic gamma-ray spectra or even sharper spectral features (like pile-ups) in TeV blazars. In this context, recently some extreme assumptions regarding the distributions of accelerated particles have been proposed. In particular, Katarzynski et al (2006) argued that a gamma-ray spectrum as hard as $\\Gamma_0 \\sim 0.7$ can be formed in a SSC model assuming a narrow parent electron distribution, e.g. power-law within $E_1$ and $E_2$, with a low-energy cutoff $E_1$ not much smaller than the high energy cutoff, $E_2$. In similar lines, Stecker et al (2007) argued that electron spectra with power law index $\\leq 1$ can be accommodated within the models of relativistic shock acceleration . It should be noted, however, that in compact objects relativistic electrons usually suffer very fast synchrotron losses, therefore the assumptions about hard electron \\textit{acceleration } cannot yet guarantee hard gamma-ray spectra. Indeed, the radiatively cooled electron spectrum cannot be harder than $dN/dE \\propto E^{-2}$, independent of the initial (acceleration) spectrum (see e.g. Aharonian 2004). If so, the inverse Compton scattering would result in a gamma-ray spectrum steeper than $E^{-1.5}$. In fact, the Klein-Nishina effect makes the spectrum even steeper. In principle, one can avoid the synchrotron cooling of electrons, e.g. in a cold ultrarelativistic wind. However such a hypothesis suggested for Mkn~501\\cite{ColdWind}, in analogy with pulsar winds, needs thorough theoretical studies to clarify whether such cold ultrarelativistic winds can be formed and survived around supermassive black holes in the cores of AGN. In this paper we suggest a new scenario which allows formation of very (in practice, arbitrary) hard gamma-ray spectra in a quite natural way. The model is based on a postulation that gamma-rays before leaving the source suffer significant photon-photon absorption due to interactions with dense radiation fields inside or in the vicinity of compact gamma-ray production region(s). Interestingly, the presence of high density radiation fields of different origin in the inner parts of blazars generally is treated as a problem for the escape of high energy gamma-radiation from their production region, and, in this regard, the current models of TeV blazars are designed in a way to avoid the internal gamma-ray absorption. Below we show that, in fact, a moderate internal photon-photon absorption can be a clue to the very hard intrinsic energy spectra of TeV blazars. ", "conclusions": "The energy spectra of VHE gamma-rays from blazars, after correction for intergalactic absorption, generally appear very hard, even for the minimum flux level of EBL determined by the integrated light of resolved galaxies at optical (Hubble) and infrared (Spitzer) wavelengths. A slight deviation from the robust lower limits of EBL leads to unusually hard intrinsic gamma-ray spectra which cannot be easily explained within the standard particle acceleration and radiation models. In this paper we suggest a scenario which can lead to the formation of instrinsic gamma-ray spectra of arbitrary hardness without introducing modifications in the particle acceleration models. The main idea is that the gamma-rays before they leave the source suffer significant internal energy-dependent absorption due to interactions with the ambient low-frequency photons. The existence of dense radiation fields of different origin in blazars (see e.g. Urry and Padovani (1995)) combined with the large photon-photon pair production cross-section, makes this scenario quite natural and effective, in particular in the compact cores of blazars. For the formation of hard VHE gamma-ray spectra, the target radiation field must have a rather narrow spectral distribution or a sharp low-energy cutoff, with a typical energy of photons of about 1 to 10 eV. Formally, for very large optical depths, this process can provide an arbitrary hardness of gamma-ray spectra, though at the expense of a significant increase of the required nonthermal energy budget. However, as long as the current blazar models require relativistically-moving gamma-ray production regions with large Doppler factors, $\\delta_j \\geq 30$, and perhaps even more \\cite{PKS_HESS,Fabian}, the available energy budget seems to be not a critical issue. The unavoidable feature of the proposed model is the radiation of secondary electrons via synchrotron or inverse Compton scattering. If the optical depth inside the gamma-ray production region is small, $\\tau \\ll 1$, e.g. the gamma-ray source is much smaller than the external source of optical photons, the secondary electrons are produced and radiate mainly outside the gamma-ray production region. Even in the case of heavy absorption of gamma-rays, the secondary radiation of secondary electrons can hardly be detected. Indeed, since the intrinsic gamma-ray luminosity is relatively modest, and the absorbed energy is re-radiated as an isotropic source\\footnote{Unless the electrons are produced in an environment with a very low magnetic field, and thus are cooled via inverse Compton scattering before any noticeable deflection.}, the lost of the beaming factor dramatically reduces the signal compared to the primary (Doppler boosted) radiation. The picure is dramatically changed when the gamma-ray source moves through a very dense photon field, such that the optical depth inside the source becomes larger than 1. In this case the main fraction of the absorbed energy is released in the form of secondary electrons inside the gamma-ray production region, and thus the radiation of the secondary electrons profits, as the primary gamma-radiation does, from the Doppler boosting. The secondary electrons are cooled through synchrotron and/or inverse Compton channels. The latter in fact proceeds via development of pair cascades as long as the typical energies of electrons or gamma-rays and the energy of target photons $\\varepsilon E_{\\rm e, \\gamma} \\gg m^2_e c^4$. The cascade, however, diminishes the energy-dependent absorption features, thus the model becomes effective when the electrons are cooled predominantly via synchrotron radiation, i.e. $B^2/8 \\pi \\geq u_{\\rm r}$. The energy density of the radiation $u_{\\rm r} = \\bar{\\varepsilon} n_{\\rm ph}$ with an average energy of target photons of about $\\bar{\\varepsilon} \\sim 1 \\ \\rm eV$ is estimated from the condition $\\tau_{\\rm max} \\geq 1$, thus for the effective suppression of the cascade \\begin{equation} B \\geq (40 \\pi \\bar{\\varepsilon}/\\sigma_{\\rm T} l)^{1/2} \\approx 0.5 (l/10^{15} \\rm cm)^{-1/2} \\tau^{1/2}_{\\rm max}\\ \\rm G \\label{B} \\end{equation} For an optical depth $\\tau_{\\rm max} \\sim 1$ and the size of the gamma-ray source $l \\sim 10^{16} \\ \\rm cm$, the magnetic field exceeding 0.1 G should be sufficient to prevent the cascade. For smaller gamma-ray production regions, e.g. $l \\sim 10^{14} \\ \\rm cm$, the magnetic field should be larger than 1 G. For such magnetic fields the synchrotron radiation of secondary electrons appears in the optical to hard X-ray energy bands. Depending on the optical depth, the synchrotron peak can be higher than the gamma-ray peak. Interestingly, unlike the classical ''synchrotron/Inverse-Compton`` models, where the ratio of the synchrotron to IC peak is determined by the ratio $u_{\\rm B}/u_{\\rm r}$, in the ''internal gamma-ray absorption`` scenario the synchrotron peak does not strongly depend on the magnetic field. Whether this scenario can be applied to the the broad-band SEDs of gamma-ray blazars, is an interesting issue which requires special dedicated studies. Finally, we want to discuss briefly the radiation mechanisms of primary gamma-radiation. Generally, the model does not give a preference to the leptonic or hadronic origin of radiation, unless the magnetic field exceeds the estimate given by Eq.(\\ref{B}). In this case the synchrotron-to-IC flux ratio produced by directly accelerated electrons would be too high, especially after the internal absorption of gamma-rays, contrary to the detected SEDs of most of the TeV blazars. Large magnetic fields in the gamma-ray production region, typically $B \\geq 1 \\ \\rm G$, would favor gamma-ray production by relativistic protons, with all advantages and disadvantages common for hadronic models. The basic problem of hadronic models is linked to the low interaction rates which do not allow the most natural explanation of the observed fast gamma-ray variability of blazars in terms of radiative cooling. For example, in the case of interactions of protons with the ambient plasma with number density $n$, the characteristic time of $pp$ interactions with production of $\\pi^0$-mesons is $t_{\\rm pp} \\approx 10^{15} n^{-1} \\ \\rm s$. Thus, in order to explain the variability of gamma-rays as short as several minutes like the TeV flares observed from PKS 21555-301 and Mkn 501, the density of plasma should be as large as $5 \\times 10^{12} \\delta_j^{-1} \\ \\rm cm^{-3}$ which implies a very heavy source and correspondingly huge kinetic energy $E_{\\rm kin}= (4/3) \\pi l^3 n m_pc^2 \\gamma_j \\approx 10^{55} \\ \\rm erg$ (here we assume that $\\delta_ \\approx \\gamma_j$). One may invoke alternative explanations of the variability of blazars, e.g. due to the adiabatic losses or escape of particles from the source, but this assumption leads to dramatic reduction of radiation efficiency, and to an increase the energy requirements to the accelerated protons. A similar problem face the photomeson processes at interactions of protons with the ambient radiation fields. Actually in the ''internal gamma-ray absorption`` scenario this mechanisms seems a quite natural choice because the same background photons which absorb gamma-rays can play a role of the target for photomeson interactions. However, because of the small cross-section, the efficiency of this process again appears quite low. The interaction time of protons with energy, $E \\geq 200 \\ \\rm MeV/(\\bar{\\varepsilon} \\gamma_j) \\simeq 2 \\times 10^{16} (\\bar{\\varepsilon}/1 \\ \\rm eV)^{-1} (\\gamma_j/10)^{-1} \\ \\rm eV$ (in the frame of the moving source with a Lorentz factor $\\gamma_j$) is estimated $t_{\\rm p \\gamma} \\approx 1/(f \\sigma_{\\rm p \\gamma} n_{\\rm ph} c) \\sim (\\sigma_{\\rm \\gamma \\gamma}/\\sigma_{\\rm p \\gamma})f^{-1} R/c \\tau_{\\rm max}^{-1}$ ($<\\sigma_{\\rm p \\gamma}> \\approx 10^{-28} \\ \\rm cm^2$ is the average cross-section and $f \\sim 0.2$ is the multiplicity of the process). Thus, we can see that during the passage of the source of optical photons of size $R$, the protons transfer only $\\sigma_{\\rm p \\gamma}/\\sigma_{\\gamma \\gamma} \\sim 10^{-3}$ fraction of their energy to gamma-rays. If such a low efficiency can be compensated by very large Doppler boosting (e.g. assuming $\\delta_j \\sim 100$), this channel can provide very large fluxes of neutrinos, which unlike gamma-rays do not suffer internal and extragalactic absorption. In the case of attenuation of VHE gamma-ray fluxes by 2 to 3 orders of magnitude, the expected fluxes of neutrinos from TeV blazars can be as large as the detection threshold of the km$^3$ volume high energy neutrino telescopes, $F_{\\nu_\\mu} (\\geq 1 \\ \\rm TeV) \\approx 10^{-11} \\ \\rm neutrinos/cm^2 s$. It is interesting to note that, because of the threshold of photomeson production, the interactions of protons of arbitrary distribution with a narrow band radiation with a characteristic energy $\\bar{\\varepsilon}$, result in a differential gamma-ray spectrum which below the energy $\\approx 10^{16} (\\bar{\\varepsilon}/1 \\ \\rm TeV)^{-1}$ eV is extremely hard, ${\\rm d}N/{\\rm d}E=\\rm const$, thus this process itself can provide very hard gamma-ray spectra independent of the spectrum of parent protons. Despite certain attractive features, this mechanism faces the same problem as $pp$ interactions - a low radiation efficiency. Therefore it can work only under conditions of extremely large Doppler boosting of radiation. The efficiency of VHE gamma-ray production can be much higher in the case of synchrotron radiation of protons, provided that the acceleration of protons proceeds at a rate close to the fundamental limit, and the magnetic field in the proton accelerator well exceeds 10 G. In particular, in the magnetic field of order $100$ G, protons can be accelerated to energies $10^{20} \\ \\rm TeV$ and thus can produce VHE synchrotron gamma-rays on timescales of $10^{4}$ s. Although due to the self-regulated synchrotron cutoff \\cite{psynch} the spectrum of gamma-rays is limited by sub-TeV energies, an observer detects Doppler boosted gamma-radiation extending to multi-TeV energies. The characteristic feature of this mechanism is the very large electromagnetic energy contained in the blob, $1/6 l^3 B^2 \\approx 2 \\times 10^{48} \\ \\rm erg$, hard X-ray emission of the secondary (pair-produced) electrons, and negligible fluxes of neutrinos. \\hyphenation{Post-Script Sprin-ger}" }, "0801/0801.4895_arXiv.txt": { "abstract": "We present the results of relic density calculations for cold dark matter candidates coming from a model of dark energy and dark matter, which is described by an asymptotically free gauge group $\\SU(2)_Z$ (QZD) with a coupling constant $\\alpha_Z \\sim$ 1 at very low scale of $\\Lambda_Z \\sim 10^{-3}$ eV while $\\alpha_Z \\sim$ weak coupling at high energies. The dark matter candidates of QZD are two fermions in the form of weakly interacting massive particles. Our results show that for masses between 50 and 285 GeV, they can account for either a considerable fraction or the entire dark matter of the Universe. ", "introduction": "\\label{sec:intro} It is almost universally accepted that the picture of the Universe made up of approximately $4\\%$ baryonic matter, $23\\%$ dark matter and $73\\%$ dark energy represents a realistic cosmological model. However, it is astounding that almost $96\\%$ of the energy density of the Universe resides in some as-yet-unknown form. What is ``dark matter''? What is ``dark energy''? In Refs.~\\cite{Hung2005,Hung2006a}, a model of dark energy and dark matter was proposed in which a new unbroken gauge group $\\SU(2)_Z$ -- the shadow sector -- grows strong at a scale $\\sim 10^{-3}\\,$eV. The gauge group $\\SU(2)_Z$ was nicknamed Quantum Zophodynamics, or QZD, in Refs.~\\cite{Hung2005,Hung2006a}, where the subscript ``\\textit{Z}\" stands for the Greek word \\textit{Zophos}, meaning darkness. The model is described by an $\\SU(2)_Z$ instanton-induced potential of an axion-like particle, $a_Z$, which possesses two degenerate minima. The degeneracy is lifted by a mechanism described in Refs.~\\cite{Hung2006a,Hung2007a}, yielding a false vacuum with energy density $\\sim (10^{-3}\\,\\text{eV})^4$ and a true vacuum with vanishing energy density. The present Universe is assumed to be trapped in the false vacuum~\\cite{Hung2007b}, whose energy density mimics the cosmological constant. This is, in a nutshell, the dark energy model proposed in Ref.~\\cite{Hung2006a}, which also computed various quantities of interest such as the tunneling rate to the true vacuum, etc. A Grand Unified Theory (GUT) involving the SM and $\\SU(2)_Z$ was considered by Ref.~\\cite{Hung2006d} (The models presented in Refs.~\\cite{Hung2006a,Hung2006d} were later revisited by Refs.~\\cite{Das}.). The particle content of the model includes two shadow fermions, $\\psi _{\\left( {L,R} \\right),i}^{\\left( Z \\right)} $ with $i=1,2$, which transform as $\\left( {1,1,0,3} \\right)$ under $\\SU(3)_c \\otimes \\SU(2)_L \\otimes \\mathrm{U}(1)_Y \\otimes \\SU(2)_Z$, two messenger scalar fields (mediating between the QZD and SM matters; one of which is much heavier than the other \\cite{Hung2006a}) $\\tilde{\\bm{\\varphi}} _i^{\\left( Z \\right)} $ with $i=1,2$ transforming as $\\left( {1,2,Y_{\\tilde \\varphi} = - 1,3} \\right)$, and one singlet complex scalar field $\\phi _Z = \\left( {1,1,0,1} \\right)$ whose imaginary part plays the role of the axion-like particle mentioned above. As discussed in Ref.~\\cite{Hung2006a}, the masses of the $\\SU(2)_Z$ triplet shadow fermions are found to be of the order of 100 - 200 GeV for the $\\SU(2)_Z$ gauge coupling to grow strong at a scale $\\sim 10^{-3}$ eV, needed for the dark energy scenario. This coupling constant starts out at GUT-scale energy with a value comparable to that of the electroweak couplings, remains relatively flat until an energy comparable to the shadow fermion masses is reached, and then starts to grow after the shadow fermions drop out of the Renormalization Group (RG) equations. At that dropout point, the $\\SU(2)_Z$ gauge coupling becomes comparable to the weak $\\SU(2)_L$ coupling at the electroweak scale energy. These features have interesting consequences concerning the possibility of the shadow fermions being candidates for cold dark matter (CDM) in the form of weakly interacting massive particles (WIMP's) \\footnote{For a review on various features of CDM and WIMP, see, e.g., Refs. \\cite{CDMRevs}.}. The main reason is the fact that the annihilation cross sections for two shadow fermions into two $\\SU(2)_Z$ ``shadow gluons'' are of the order of the weak cross sections, a typical requirement for WIMP's. The estimates that were made in Ref.~\\cite{Hung2006a} showed that it was possible for shadow fermions to be candidates for CDM with the \\textit{right} relic density. In this work, we would like to investigate this scenario in more details and by solving shadow fermions' evolution equations to determine the conditions under which they can be considered to be WIMP cold dark matter candidates. It will be seen that the mass range for the shadow fermions obtained by the requirement of having the \\textit{right} density fits in snugly with that used in the RG equations (i.e., the $\\SU(2)_Z$ gauge coupling grows strong at a scale $\\sim 10^{-3}$ eV). The outline of the paper is as follows. First, we go over the QZD model as far as the issue of dark matter is concerned. Then, we derive the evolution equations for shadow fermions and consequently solve them numerically, to obtain their relic density. Finally, the results of our relic density calculations will be presented and discussed, in comparison with the observational values. The shadow fermions relic density, when computed, would only depend on their masses. Therefore, the parameter space is simply two dimensional. ", "conclusions": "We solved evolution equations for number densities of shadow fermions and obtained their total present-day density. The heavier shadow fermion turned out to be long lived if its mass differs from that of the messenger field. In that case, our results revealed an upper bound on the mass of the heavier shadow fermion, i.e., $m_2 \\approx 245$ GeV, above which its \\textit{late} decay can potentially disturb the CMB density of the Universe beyond the measured fluctuation level of $10^{-5}$. For lighter shadow fermions, the total relic density can account for the entire dark matter of the Universe depending on the mass combination of shadow fermions. When the total density falls short of the observationally suggested density, it still, for most of masses, provides significant fraction of the dark matter of the Universe. Our results showed that if the heavier shadow fermion's mass is large, considerable mass differences would be needed to comply with experimental bounds. On the other hand, if the heavier shadow fermion's mass is small, little or even no mass differences suffice to give the right relic density. In that sense, degenerate and near-degenerate mass cases become relevant at low mass scales, but not for less than 50 GeV. A very short lifetime is expected for the heavier shadow fermion if its mass is the same as that of the messenger field. In that case, the calculations reduce to a one-species case. Our results suggest that a sole shadow fermion must have a mass of about 190~--~210 GeV to account for the whole dark matter of the Universe. Last but not least, possible detections of the shadow fermion CDM candidates are briefly discussed in Ref.~\\cite{Hung2006a}. Needless to say, more work along this line is warranted for this model." }, "0801/0801.0116.txt": { "abstract": "Primordial black holes (PBHs) are a profound signature of primordial cosmological structures and provide a theoretical tool to study nontrivial physics of the early Universe. The mechanisms of PBH formation are discussed and observational constraints on the PBH spectrum, or effects of PBH evaporation, are shown to restrict a wide range of particle physics models, predicting an enhancement of the ultraviolet part of the spectrum of density perturbations, early dust-like stages, first order phase transitions and stages of superheavy metastable particle dominance in the early Universe. The mechanism of closed wall contraction can lead, in the inflationary Universe, to a new approach to galaxy formation, involving primordial clouds of massive BHs created around the intermediate mass or supermassive BH and playing the role of galactic seeds. Primordial black holes (PBHs) are a profound signature of primordial cosmological structures and provide a theoretical tool to study nontrivial physics of the early Universe. The mechanisms of PBH formation are discussed and observational constraints on the PBH spectrum, or effects of PBH evaporation, are shown to restrict a wide range of particle physics models, predicting an enhancement of the ultraviolet part of the spectrum of density perturbations, early dust-like stages, first order phase transitions and stages of superheavy metastable particle dominance in the early Universe. The mechanism of closed wall contraction can lead, in the inflationary Universe, to a new approach to galaxy formation, involving primordial clouds of massive BHs created around the intermediate mass or supermassive BH and playing the role of galactic seeds. ", "introduction": "The convergence of the frontiers of our knowledge in micro- and macro- worlds leads to the wrong circle of problems, illustrated by the mystical Uhroboros (self-eating-snake). The Uhroboros puzzle may be formulated as follows: {\\it The theory of the Universe is based on the predictions of particle theory, that need cosmology for their test}. Cosmoparticle physics \\cite{ADS,MKH,book,book3,Khlopov:2004jb,bled} offers the way out of this wrong circle. It studies the fundamental basis and mutual relationship between micro-and macro-worlds in the proper combination of physical, astrophysical and cosmological signatures. Some aspects of this relationship, which arise in the astrophysical problem of Primordial Black Holes (PBH) is the subject of this review. In particle theory Noether's theorem relates the exact symmetry to conservation of respective charge. Extensions of the standard model imply new symmetries and new particle states. The respective symmetry breaking induces new fundamental physical scales in particle theory. If the symmetry is strict, its existence implies new conserved charge. The lightest particle, bearing this charge, is stable. It gives rise to the deep relationship between dark matter candidates and particle symmetry beyond the Standard model. The mechanism of spontaneous breaking of particle symmetry also has cosmological impact. Heating of condensed matter leads to restoration of its symmetry. When the heated matter cools down, phase transition to the phase of broken symmetry takes place. In the course of the phase transitions, corresponding to given type of symmetry breaking, topological defects can form. One can directly observe formation of such defects in liquid crystals or in superfluid He. In the same manner the mechanism of spontaneous breaking of particle symmetry implies restoration of the underlying symmetry. When temperature decreases in the course of cosmological expansion, transitions to the phase of broken symmetry can lead, depending on the symmetry breaking pattern, to formation of topological defects in very early Universe. Defects can represent new forms of stable particles (as it is in the case of magnetic monopoles \\cite{t'Hooft,polyakov,kz,Priroda,preskill,SAOmonop}), or extended structures, such as cosmic strings \\cite{zv1,zv2} or cosmic walls \\cite{okun}. In the old Big bang scenario cosmological expansion and its initial conditions were given {\\it a priori} \\cite{Weinberg,ZNSEU}. In the modern cosmology expansion of Universe and its initial conditions are related to inflation \\cite{Star80,Guthinfl,Linde:1981mu,Albrecht,Linde:1983gd}, baryosynthesis and nonbaryonic dark matter (see review in \\cite{Lindebook,Kolbbook}). Physics, underlying inflation, baryosynthesis and dark matter, is referred to extensions of the standard model, and variety of such extensions makes the whole picture in general ambiguous. However, in a framework of each particular physical realization of inflationary model with baryosynthesis and dark matter the corresponding model dependent cosmological scenario can be specified in all details. In such scenario main stages of cosmological evolution, structure and physical content of the Universe reflect structure of the underlying physical model. The latter should include with necessity the standard model, describing properties of baryonic matter, and its extensions, responsible for inflation, baryosynthesis and dark matter. In no case cosmological impact of such extensions is reduced to reproduction of these three phenomena only. A nontrivial path of cosmological evolution, specific for each particular realization of inflational model with baryosynthesis and nonbaryonic dark matter, always contains some additional model dependent cosmologically viable predictions, which can be confronted with astrophysical data. Here we concentrate on Primordial Black Holes as profound signature of such phenomena. It was probably Pierre-Simon Laplace \\cite{Laplace} in the beginning of XIX century, who noted first that in very massive stars escape velocity can exceed the speed of light and light can not come from such stars. This conclusion made in the framework of Newton mechanics and Newton corpuscular theory of light has further transformed into the notion of \"black hole\" in the framework of general relativity and electromagnetic theory. Any object of mass $M$ can become a black hole, being put within its gravitational radius $r_g=2 G M/c^2.$ At present time black holes (BH) can be created only by a gravitational collapse of compact objects with mass more than about three Solar mass \\cite{1,ZNRA}. It can be a natural end of massive stars or can result from evolution of dense stellar clusters. However in the early Universe there were no limits on the mass of BH. Ya.B. Zeldovich and I.D. Novikov \\cite{ZN} noticed that if cosmological expansion stops in some region, black hole can be formed in this region within the cosmological horizon. It corresponds to strong deviation from general expansion and reflects strong inhomogeneity in the early Universe. There are several mechanisms for such strong inhomogeneity and we'll trace their links to cosmological consequences of particle theory. %The simplest one is a collapse of strongly inhomogeneous regions %just after the end of inflation (see e.g. \\cite{2}). Another %possible source of BH could be a collapse of cosmic strings \\cite{3} %that are produced in early phase transitions with symmetry breaking. %The collisions of the bubble walls \\cite{4,5} created at phase %transitions of the first order can lead to a primordial black hole %(PBH) formation. Primordial Black Holes (PBHs) are a very sensitive cosmological probe for physics phenomena occurring in the early Universe. They could be formed by many different mechanisms, {\\it e.g.}, initial density inhomogeneities \\cite{hawking1,hawkingCarr} and non-linear metric perturbations \\cite{Bullock:1996at,Ivanov:1997ia,Bullock:1998mi}, blue spectra of density fluctuations \\cite{Khlopov:1984wc,polnarev,Lidsey:1995ir,Kotok:1998rp, Dubrovich02,Sendouda:2006nu}, a softening of the equation of state \\cite{canuto,Khlopov:1984wc,polnarev}, development of gravitational instability on early dust-like stages of dominance of supermassive particles and scalar fields \\cite{khlopov0,polnarev0,polnarev1,khlopov1} and evolution of gravitationally bound objects formed at these stages \\cite{Kalashnikov,Kadnikov}, collapse of cosmic strings \\cite{hawking2,Polnarev:1988dh,Hansen:2000jv,Cheng:1996du,Nagasawa:2005hv} and necklaces \\cite{Matsuda:2005ez}, a double inflation scenario \\cite{nas,Kim:1999xg,Yamaguchi:2001zh,Yamaguchi:2002sp}, first order phase transitions \\cite{hawking3,pt1Jedamzik,kkrs,kkrs1,kkrs2}, a step in the power spectrum \\cite{Sakharov0,polarski1}, etc. (see \\cite{polnarev,book,book3,Carr:2003bj,book2} for a review). Being formed, PBHs should retain in the Universe and, if survive to the present time, represent a specific form of dark matter \\cite{khlopov7,Ivanov:1994pa,book,book3,Blais:2002nd,Chavda:2002cj, Afshordi:2003zb,book2,Chen:2004ft}. Effect of PBH evaporation by S.W.Hawking \\cite{hawking4} makes evaporating PBHs a source of fluxes of products of evaporation, particularly of $\\gamma$ radiation \\cite{Page:1976wx}. MiniPBHs with mass below $10^{14}$~g evaporate completely and do not survive to the present time. However, effect of their evaporation should cause influence on physical processes in the early Universe, thus providing a test for their existence by methods of cosmoarcheology \\cite{Cosmoarcheology}, studying cosmological imprints of new physics in astrophysical data. In a wide range of parameters the predicted effect of PBHs contradicts the data and it puts restrictions on mechanism of PBH formation and the underlying physics of very early Universe. On the other hand, at some fixed values of parameters, PBHs or effects of their evaporation can provide a nontrivial solution for astrophysical problems. Various aspects of PBH physics, mechanisms of their formation, evolution and effects are discussed in \\cite{carr1,carrMG,LGreen,khlopov6,polnarev,Grillo:1980uj, Chapline:1975tn,Hayward:1989jq,Yokoyama:1995ex,Kim:1996hr, Heckler:1997jv,MacGibbon:2007yq,Page:2007yr,green,Niemeyer:1997mt,Kribs:1999bs, Green:2004wb,Yokoyama:1998pt,Yokoyama:1998xd,Yokoyama:1999xi, Bringmann:2001yp,Dimopoulos:2003ce,Nozari:2007kv,LythMalik,Zaballa:2006kh, Harada:2004pf, Custodio:2005en,Bousso:1995cc,Bousso:1996wy,Elizalde:1999dw, Nojiri:1999vv,Bousso:1999iq,Silk:2000em,Polarski:2001jk, Barrow:1996jk,Paul:2000jb,Paul:2001yt,Paul:2005bk,Polarski:2001yn, Carr:1993aq,Yokoyama:1998qw,Kaloper:2004yj,Pelliccia:2007fh, Stojkovic:2005zh,Soda,Ahn:2006uc,babichev4,babichev5,babichev6,Guariento:2007bs} particularly specifying PBH formation and effects in braneworld cosmology \\cite{Guedens:2002km,Guedens:2002sd,Clancy:2003zd, Tikhomirov:2005bt}, on inflationary preheating \\cite{Bassett:2000ha}, formation of PBHs in QCD phase transition \\cite{Jedamzik:1998hc, Widerin:1998my}, properties of superhorizon BHs \\cite{Harada:2005sc,Harada:2006gn}, role of PBHs in baryosynthesis \\cite{Grillo:1980rt,Barrow:1990he,Turner:1979bt,Upadhyay:1999vk, Bugaev:2001xr}, effects of PBH evaporation in the early Universe and in modern cosmic ray, neutrino and gamma fluxes \\cite{mujana,Fegan:1978zn,Green:2001kw,Frampton:2005fk, MacGibbon:1990zk,MacGibbon:1991tj,Halzen:1991uw,Halzen:1995hu, Bugaev:2000bz,Bugaev:2002yt,Volkova:1994fb,Gibilisco:1996ft, Golubkov:2000qy,He:2002vz,Gibilisco:1996dk,Custodio:2002jv, Sendouda:2003dc,Maki:1995pa,barraupbar,Barrau:2002mc, Wells:1998jv,Cline:1996uk,Xu:1998hn,Cline:1998fx,Sendouda:2006yc, Barrau:1999sk,Derishev:1999xn,Tikhomirov:2004rs,Seto,barrau, barraugamma,barrauprd}, in creation of hypothetical particles \\cite{Bell:1998jk,lemoine,green1,barrau2}, PBH clustering and creation of supermassive BHs \\cite{Bean:2002kx,Duechting:2004dk,Chisholm:2005vm, Dokuchaev:2004kr,Mack:2006gz,Rubin:2005pq}, effects in cosmic rays and colliders from PBHs in low scale gravity models \\cite{barrauADDBH,barrauADDac}. Here we outline the role of PBHs as a link in cosmoarcheoLOGICAL chain, connecting cosmological predictions of particle theory with observational data. We discuss the way, in which spectrum of PBHs reflects properties of superheavy metastable particles and of phase transitions on inflationary and post-inflationary stages. We briefly review possible cosmological reflections of particle physics (section \\ref{Cosmophenomenology}), illustrate in section \\ref{dust} some mechanisms of PBH formation on stage of dominance of superheavy particles and fields (subsection \\ref{particles}) and from second order phase transition on inflationary stage. Effective mechanism of BH formation during bubble nucleation provides a sensitive tool to probe existence of cosmological first order phase transitions by PBHs (section \\ref{phasetransitions}). Existence of stable remnants of PBH evaporation can strongly increase the sensitivity of such probe and we demonstrate this possibility in section \\ref{gravitino} on an example of gravitino production in PBH evaporation. Being formed within cosmological horizon, PBHs seem to have masses much less than the mass of stars, constrained by small size of horizon in very early Universe. However, if phase transition takes place on inflationary stage, closed walls of practically any size can be formed (subsection \\ref{walls}) and their successive collapse can give rise to clouds of massive black holes, which can play the role of seeds for galaxies (section \\ref{MBHwalls}). The impact of constraints and cosmological scenarios, involving primordial black holes, is briefly discussed in section \\ref{Discussion}. ", "conclusions": "\\label{Discussion} For long time scenarios with Primordial Black holes belonged dominantly to cosmological {\\it anti-Utopias}, to \"fantasies\", which provided restrictions on physics of very early Universe from contradiction of their predictions with observational data. Even this \"negative\" type of information makes PBHs an important theoretical tool. Being formed in the very early Universe as initially nonrelativistic form of matter, PBHs should have increased their contribution to the total density during RD stage of expansion, while effect of PBH evaporation should have strongly increased the sensitivity of astrophysical data to their presence. It links astrophysical constraints on hypothetical sources of cosmic rays or gamma background, on hypothetical factors, causing influence on light element abundance and spectrum of CMB, to restrictions on superheavy particles in early Universe and on first and second order phase transitions, thus making a sensitive astrophysical probe to particle symmetry structure and pattern of its breaking at superhigh energy scales. Gravitational mechanism of particle creation in PBH evaporation makes evaporating PBH an unique source of any species of particles, which can exist in our space-time. At least theoretically, PBHs can be treated as source of such particles, which are strongly suppressed in any other astrophysical mechanism of particle production, either due to a very large mass of these species, or owing to their superweak interaction with ordinary matter. By construction astrophysical constraint excludes effect, predicted to be larger, than observed. At the edge such constraint converts into an alternative mechanism for the observed phenomenon. At some fixed values of parameters, PBH spectrum can play a positive role and shed new light on the old astrophysical problems. The common sense is to think that PBHs should have small sub-stellar mass. Formation of PBHs within cosmological horizon, which was very small in very early Universe, seem to argue for this viewpoint. However, phase transitions on inflationary stage can provide spikes in spectrum of fluctuations at any scale, or provide formation of closed massive domain walls of any size. In the latter case primordial clouds of massive black holes around intermediate mass or supermassive black hole is possible. Such clouds have a fractal spatial distribution. A development of this approach gives ground for a principally new scenario of the galaxy formation in the model of the Big Bang Universe. Traditionally, Big Bang model assumes a homogeneous distribution of matter on all scales, whereas the appearance of observed inhomogeneities is related to the growth of small initial density perturbations. However, the analysis of the cosmological consequences of the particle theory indicates the possible existence of strongly inhomogeneous primordial structures in the distribution of both the dark matter and baryons. These primordial structures represent a new factor in galaxy formation theory. Topological defects such as the cosmological walls and filaments, primordial black holes, archioles in the models of axionic CDM, and essentially inhomogeneous baryosynthesis (leading to the formation of antimatter domains in the baryon-asymmetric Universe \\cite{exl1,exl2,crg,kolb,we,khl,CSKZ,zil,sb,dolgmain, khlopgolubkov,bgk,FarKhl,book,book2,book3}) offer by no means a complete list of possible primary inhomogeneities inferred from the existing elementary particle models. Observational cosmology offers strong evidences favoring the existence of processes, determined by new physics, and the experimental physics approaches to their investigation. Cosmoparticle physics \\cite{ADS,MKH,book,book3}, studying the physical, astrophysical and cosmological impact of new laws of Nature, explores the new forms of matter and their physical properties. Its development offers the great challenge for theoretical and experimental research. Physics of Primordial Black holes can play important role in this process." }, "0801/0801.2157_arXiv.txt": { "abstract": "{ Inflation may occur while rolling into the metastable supersymmetry-breaking vacuum of massive supersymmetric QCD. We explore the range of parameters in which slow-roll inflation and long-lived metastable supersymmetry breaking may be simultaneously realized. The end of slow-roll inflation in this context coincides with the spontaneous breaking of a global symmetry, which may give rise to significant curvature perturbations via inhomogenous preheating. Such spontaneous symmetry breaking at the end of inflation may give rise to observable non-gaussianities, distinguishing this scenario from more conventional models of supersymmetric hybrid inflation.} \\begin{document} ", "introduction": "Spontaneously broken supersymmetry (SUSY) has long been one of the most attractive possibilities for physics beyond the Standard Model \\cite{Dimopoulos:1981zb}. Considerable effort has been devoted to elucidating dynamical mechanisms of SUSY-breaking, in which strong-coupling effects give rise to a naturally small SUSY-breaking scale \\cite{Witten:1981nf}. The SUSY-breaking vacua of such theories need not be global minima of the potential, and metastability of phenomenologically-viable vacua appears to be generic when embedding the MSSM into a larger setting. Indeed, the study of metastable dynamical SUSY-breaking has recently undergone something of a renaissance, catalyzed by the observation that massive supersymmetric QCD (SQCD) possesses SUSY-breaking local minima whose lifetimes can be longer than the present age of the universe \\cite{Intriligator:2006dd} (for an excellent review, see \\cite{Intriligator:2007cp}). The simplicity and genericness of such theories makes them particularly well-suited to supersymmetric model-building. Far from serving only as a possible resolution of the hierarchy problem, spontaneously-broken supersymmetry emerges frequently in inflationary settings as well \\cite{Guth:1980zm, Linde:1981mu, Albrecht:1982wi, Linde:1983gd}. There exist a plethora of possible models aimed at realizing inflation in a natural context \\cite{Lyth:1998xn}, many of which exploit supersymmetry and supersymmetry breaking \\cite{Kinney:1997hm, Kinney:1998dv, Dvali:1994ms, Dimopoulos:1997fv, Copeland:1994vg}. Based on the moderate success of such theories, it is tempting to suppose that inflation may be built into a supersymmetry-breaking sector. Indeed, it has already been demonstrated that inflation may arise in simple strongly-coupled gauge theories \\cite{Dimopoulos:1997fv}. This suggests that it may be possible to inflate while rolling into the metastable vacuum of massive SQCD from sub-Planckian initial field values. The appeal of successful inflation in this scenario is twofold: first, it provides a generic and technically natural means of simultaneously realizing inflation and supersymmetry breaking; second, it furnishes a robust UV completion for the inflationary sector. Moreover, the spontaneous global symmetry-breaking that accompanies the end of inflation in these models may give rise to additional curvature perturbations exhibiting observable non-gaussianities. In this note we address the prospects for, and signatures of, slow-roll inflation in models with metastable dynamical supersymmetry breaking. The organization of the paper is as follows. In Sec.~\\ref{sec:hybrid} we review the traditional scenario of supersymmetric hybrid inflation proposed in \\cite{Dvali:1994ms}, wherein one-loop corrections drive inflation along a flat direction of a theory with a relatively simple superpotential. In Sec.~\\ref{sec:hybridSQCD} we review the means by which supersymmetric hybrid inflation may be realized in strongly-coupled gauge theories \\cite{Dimopoulos:1997fv}. Subsequently, in Sec.~\\ref{sec:ISS} we turn to the Intriligator, Seiberg, and Shih (ISS) model \\cite{Intriligator:2006dd} of metastable supersymmetry breaking in supersymmetric QCD and explore its implications. In the vicinity of the supersymmetry breaking metastable vacuum, features of the ISS model are reminiscent of supersymmetric hybrid inflation. With this motivation, in Sec.~\\ref{sec:metastable} we explore the prospects for, and constraints on, inflation while rolling into the metastable supersymmetry breaking vacuum of the ISS model; in Sec.~\\ref{sec:predictions} we discuss the subsequent inflationary predictions assuming the inflaton is primarily responsible for the primordial curvature perturbation. However, the spontaneous symmetry breaking that accompanies the end of slow-roll inflation may give rise to additional curvature perturbations through inhomogenous preheating \\cite{Kolb:2004jm}. In Sec.~\\ref{sec:broken} we explore this possibility and its consequences, including the prospects for observable non-gaussianities. In Sec.~\\ref{sec:conclusions} we conclude the analysis and consider directions for further work. ", "conclusions": "\\label{sec:conclusions} We have seen that inflation may naturally occur while rolling into the supersymmetry-breaking metastable vacuum of massive supersymmetric QCD. Although the combination of supersymmetry-breaking and inflation is more a matter of novelty than profundity, this scenario is particularly attractive in that it contains a concrete UV completion of the inflationary sector and the potential for distinctive observational signatures. Successful slow-roll inflation in the ISS model requires a large number of flavors and relatively small magnetic gauge symmetry, as well as a natural hierarchy $m \\ll \\mu \\ll \\Lambda \\lsim M_P.$ Although quadratic corrections to the inflationary potential arise in the presence of a non-canonical K\\\"{a}hler potential, the ensuing contribution to slow-roll parameters may be small enough to forestall the conventional SUSY $\\eta$ problem. Moreover, the spontaneous global symmetry breaking that accompanies the end of slow-roll inflation in the ISS model may give rise to dominant curvature perturbations through inhomogenous preheating. Such perturbations may possess observable non-gaussianities, further distinguishing ISS-flation from its more conventional cousins. It is difficult to simultaneously obtain weak-scale SUSY breaking and the observed inflationary spectrum strictly from the dynamics of the ISS model; standalone ISS-flation would seem to favor split SUSY or other high-scale mediation. However, weak-scale SUSY-breaking using conventional gauge- or gravity-mediation is feasible if the primary contribution to primordial curvature perturbations arises from coupling to a separate preheating sector. It would be interesting to consider concrete realizations of split supersymmetry using ISS SUSY-breaking dynamics, given the high scale of SUSY-breaking required to match inflationary observables in the absence of inhomogenous preheating. Likewise, a more careful analysis of non-gaussianities arising from inhomogenous preheating may be useful in light of recent evidence for significant non-gaussianities in the WMAP 3-year data \\cite{Yadav:2007yy}." }, "0801/0801.2966_arXiv.txt": { "abstract": "{ Known classes of radio wavelength transients range from the nearby---stellar flares and radio pulsars---to the distant Universe---$\\gamma$-ray burst afterglows. Hypothesized classes of radio transients include analogs of known objects, e.g., extrasolar planets emitting Jovian-like radio bursts and giant-pulse emitting pulsars in other galaxies, to the exotic, prompt emission from $\\gamma$-ray bursts, evaporating black holes, and transmitters from other civilizations. A number of instruments and facilities are either under construction or in early observational stages and are slated to become available in the next few years. With a combination of wide fields of view and wavelength agility, the detection and study of radio transients will improve immensely.} ", "introduction": " ", "conclusions": "" }, "0801/0801.2820_arXiv.txt": { "abstract": "High spatial resolution spectro-polarimetric observation of a decaying spot was observed with the Diffraction Limited Spectro-Polarimeter. The spatial resolution achieved was close to the diffraction limit (0.\"18) of the Dunn Solar Telescope. The fine scales present inside the decaying active region as well as surrounding areas were studied. Two interesting phenomenon observed are: (i) Canopy like structures are likely to be present in the umbral dots as well as in the light bridges providing evidence for field-free intrusion, (ii) There are opposite polarity loops present outside of the spot and some of them connects to the main spot and the surrounding magnetic features. ", "introduction": "\\label{sec:intro} Over the past few years, high spatial resolution solar observations became feasible with the development of versatile Adaptive Optics (AO; Rimmele, 2004) systems. The success in obtaining consistent high spatial resolution images, from the ground, revived the development of new instrumentation for observations close to the telescopes' diffraction limit. The versatility of the solar instrumentation made it possible to simultaneously observe a field-of-view (FOV) of interest at different wavelengths. Simultaneous imaging and spectroscopic observations are feasible for quantitative study of physical parameters. The success of the AO system is well appreciated during spectroscopic and spectro-polarimetric observations. The long integration time and the time required to scan the FOV of interest imposes the consistent image quality requirement from the AO for longer durations (as long as an hour). To supplement the spectroscopic observations, a set of imageries are also essential in order to track the features of interest. One such combination was developed for the Dunn Solar Telescope (DST) of the National Solar Observatory (NSO) at Sacramento Peak, Sunspot situated in New Mexico, USA. An exit port was assigned as a dedicated port for fixed and well calibrated instrumentation where several set of instruments were deployed and facilitate observations with very minimal setup time. This setup also facilitates standardised data reduction procedures. A set of instrumentation, G-band \\& Ca-K imagery, Diffraction Limited Spectro-polarimeter (DLSP), and a light feed for either a Universal Birefringent Filter (UBF) setup or a dual Fabry-Perot setup, were deployed in this port. More details on this instrumentation can be found in Sankarasubramanian et al. (2004; 2006) and Rimmele \\& Sankarasubramanian (2004). The diffraction limited imaging capability of the current day telescopes produced several recent papers on the fine scale structures of active as well as quiet regions. Langhans et al. (2005), Lites \\& Socas-Navarro (2004), Sankarasubramanian \\& Rimmele (2003), Sankarasubramanian, Rimmele, \\& Lites (2004), Rimmele \\& Sankarasubramanian (2004), and Rimmele \\& Marino (2006) are few examples from a long list. The positive aspects of some of these high resolution observations are the availability of vector magnetic field data close to the diffraction limit and the provision to observe Doppler velocities and other physical parameters near simultaneously. With several available space programs, it is now even possible to co-ordinate ground based observations with space borne instruments. In this paper, we report on an observation of a decaying active region using the DST and the fixed port instrumentation. We used the DLSP along with G-band and Ca-K imagers to study the vector magnetic field and its morphology of the decaying active. H-alpha images were also simultaneously obtained using the UBF. These observations were also co-ordinated with TRACE observations, however the TRACE observations will not be discussed in this paper. Section 2 discusses the observing setup and details of the observations. Preliminary results from these observations are spelled out in section 3. ", "conclusions": "High spatial resolution observation of a decaying active region NOAA 0781 were studied. A set of instruments, DLSP, G-band \\& CaK imageries, and UBF for H-alpha images were used to obtain the observations. The data were then analysed for the morphological structures of the small scale fields present in and around this active region. The variations of the Stokes profiles observed in and around the UD and LB of this active region support the field free intrusion model suggested by Parker (1979). It is also observed that the edges of the LBs show strong linear polarisation signals compared to the center. The small scale fields surrounding this active region show opposite polarity profiles connected to the parent spot as well as to the surrounding magnetic field regions. This active region was also monitored using the Michelson Doppler Imager (MDI) on-board SoHO. The spot decayed into plages few days after our observations. The MDI movie also showed bunch of opposite polarity moving magnetic features surging out of the parent spot. The opposite polarity small scale fields observed in the DLSP may belong to one such field region during the surge. A more detailed analysis combining the timing of the MDI observations with the DLSP observation is required to confirm this. The data from the co-ordinated TRACE observations may provide some more evidence on the morphological structures of these small scale loops at higher atmospheric layers." }, "0801/0801.4017_arXiv.txt": { "abstract": "We present multi-waveband optical imaging data obtained from observations of the Subaru/XMM-Newton Deep Survey (SXDS). The survey field, centered at R.A. = $02^{h}18^{m}00^{s}$, decl. = $-05^{\\circ}00'00''$, has been the focus of a wide range of multi-wavelength observing programs spanning from X-ray to radio wavelengths. A large part of the optical imaging observations are carried out with Suprime-Cam on Subaru Telescope at Mauna Kea in the course of Subaru Telescope ``Observatory Projects''. This paper describes our optical observations, data reduction and analysis procedures employed, and the characteristics of the data products. A total area of 1.22 deg$^{2}$ is covered in five contiguous sub-fields, each of which corresponds to a single Suprime-Cam field of view ($\\sim34'\\times 27'$ ), in five broad-band filters $B, V, R_c, i', z'$ to the depths of $B=28.4, V=27.8, R_c=27.7, i'=27.7$ and $z'=26.6$ ($AB, 3\\,\\sigma, \\phi=2''$). The data are reduced and compiled into five multi-waveband photometric catalogs, separately for each Suprime-Cam pointing. The $i'$-band catalogs contain about 900,000 objects, making the SXDS catalogs one of the largest multi-waveband catalogs in corresponding depth and area coverage. The SXDS catalogs can be used for an extensive range of astronomical applications such as the number density of the Galactic halo stars to the large scale structures at the distant universe. The number counts of galaxies are derived and compared with those of existing deep extragalactic surveys. The optical data, the source catalogs, and configuration files used to create the catalogs are publicly available via the SXDS web page ({\\tt http://www.naoj.org/Science/SubaruProject/SXDS/index.html}). ", "introduction": "\\label{sec:intro} % Understanding the formation and evolution processes of the individual galaxy and the growth of the large scale structures (LSSs) in the universe is one of the major goals in extragalactic astronomy today. In the scheme of a typical $\\Lambda$-CDM model, the growth of structures is governed by the gravitational growth of initial fluctuations of dark matter. The baryonic material cools in these dark matter structures and grows through hierarchical clustering to galaxies and clusters of galaxies. The subsequent evolution of galaxies is closely connected to their environments, which, of course, relate to the LSSs, where those galaxies are located. The ultimate goal of extragalactic surveys is to trace these evolutionary processes by well-defined statistical galaxy samples. Optical imaging is arguably the cornerstone of any extragalactic survey, since it provides identifications and positions of celestial objects for follow-up spectroscopy. In recent years, there have been a number of deep imaging survey projects which devote significant amounts of telescope time, such as Great Observatories Origins Deep Survey (GOODS: Giavalisco et al. 2004) and Cosmic Evolution Survey (COSMOS: HST treasury project: Feldmann et al. 2006; Scoville et al. 2007). These surveys provide multi-waveband galaxy samples at faint magnitude. Then, the photometric redshift techniques (Furusawa et al. 2000; Bolzonella et al. 2000; Feldmann et al. 2006) are frequently used to pre-select candidates for spectroscopy. Although a great deal of information can be obtained from the optical data alone, the value of the data set grows significantly as data at other wavelengths from other facilities are added. We therefore elected to use a multi-wavelength approach for the Subaru/XMM-Newton Deep Survey (SXDS; Sekiguchi et al. 2007, in preparation) from the very start, ensuring that our chosen field would be accessible and suitable for observations at all wavelengths. An equatorial field, centered at R.A. = $02^{h}18^{m}00^{s}$, decl. = $-05^{\\circ}00'00''$ is chosen to tie up with the deep X-ray observations with XMM-Newton observatory (Ueda et al. 2007), and also due to the accessibility by all major observatories, both existing and planned. The major multi-wavelength observation programs on the SXDS field (hereafter SXDF) include deep radio imaging with the VLA (Simpson et al. 2006), sub-mm mapping with SCUBA (Coppin et al. 2006; Mortier et al., 2005), mid-infrared observations with Spitzer Space Telescope (Lonsdale et al. 2003), deep near-infrared imaging with the UKIRT/WFCAM (Foucaud et al. 2006; Dye et al. 2006; Lawrence et al. 2007), and the X-ray observations with XMM-Newton observatory. Importantly, the survey field of an infrared ultra deep survey (UDS; Foucaud et al. 2006) covering 0.77 square degrees as part of the UKIDSS project (Lawrence et al. 2007) is centered on the SXDF. It is expected that our extensive multi-wavelength data set will provide photometric redshifts accurate to $\\Delta z\\lesssim 0.1$ over a wide range of redshift, as well as detailed spectral energy distributions for the vast majority of the objects in the field. The SXDS has been undertaken as a part of the Subaru Telescope ``Observatory Projects'', in which a large amount of observing times are devoted to carry out intensive survey programs by combining observing times rewarded to builders and observatory staff of the Subaru Telescope. Note that Subaru Deep Field (SDF; Kashikawa et al. 2004) is another observatory project, which targets on a field different from the SXDF. The data set of the SDF is slightly deeper than the SXDF, but concentrates on one-fifth narrower field than the SXDF. Thus, these surveys complement each other and they can be used for a wide variety of studies. In this paper, we describe the observation details and data reduction and analysis of our optical imaging survey with the prime-focus camera on Subaru Telescope, as well as the resultant data products. A detailed description of the survey strategies, scientific objectives, and multi-wavelength survey plans is given in a companion paper (Sekiguchi et al., 2007 - Paper I, in preparation). In Section~2, we describe optical imaging observations by Subaru Telescope, and in Section~3 we explain the data reduction procedure employed. We present creation of multi-waveband photometric catalogs including object detection and aperture photometry in Section~4, and calibrations of photometry and astrometry in Section~5. In Section~6, we discuss characteristics of the catalogs such as limiting magnitude, detection completeness, and number counts of galaxies. The summary is given in Section~7. Throughout this paper, all the magnitudes are expressed in the AB system unless otherwise mentioned. ", "conclusions": "% The optical imaging observations of the SXDS project are carried out using the Suprime-Cam on Subaru Telescope in $B, V, R_c, i'$ and $z'$ bands. The SXDF has a contiguous area coverage of $\\sim$1.2 deg$^{2}$, which consists of five Suprime-Cam pointings. The photometric zeropoints are determined with an absolute accuracy of no larger than 0.05 mag r.m.s. in the photometry. The r.m.s. of the astrometric accuracies across the field is on the order of 0.2 arcsec in both RA and Dec. The systematic differences in the adopted world coordinates between different pointings are within about 1 arcsec at the outer edge of the field of view. Thus, our SXDS catalogs have a good enough positional accuracies to perform follow-up spectroscopy. Multi-waveband photometric catalogs of detected objects are created for each band and for each pointing. Each of the catalogs contains more than a hundred sixty thousand of objects. The catalogs have quite homogeneous S/Ns across the field. The achieved limiting magnitudes in each band are $B=28.4, V=27.8, R_{c}=27.7, i'=27.7$ and $z'=26.6$ ($AB, 3\\,\\sigma, \\phi = 2''$). The detection completeness as a function of magnitude is estimated by a Monte-Carlo simulation assuming a Gaussian profile. The number counts of galaxies in the SXDF in each band are computed by correcting the detection completeness, which are consistent with each other among the five pointings with a slight difference at faint magnitudes. The mean number counts of galaxies averaged over the five pointings show a good agreement with results from previous surveys down to the faint-end magnitude. With the aid of the wide coverage area of SXDS data, the uncertainty of the number counts of galaxies due to the Poisson error is greatly decreased. It is emphasized that the SXDS data is extremely useful for pursuing studies on celestial objects spreading in a wide field without suffering from the Poisson error and the field-to-field variation. The SXDS catalogs can be applied to studies, scientific objectives of which range from the Galactic objects to the large scale structures of the universe. The optical data, the compiled photometric catalogs, and configuration files used to create the catalogs have been released to the public and can be retrieved from the public data archives server of National Astronomical Observatory of Japan, {\\tt http://www.naoj.org/Science/SubaruProject/SXDS/index.html}. \\smallskip Acknowledgments We thank Dr. Masafumi Yagi for his kind suggestions in analyzing the data based on the software package developed by himself. We thank Dr. Nobunari Kashikawa for fruitful discussions. The anonymous referee would also be appreciated for his/her constructive suggestions. We would like to convey our gratitude to the Subaru Telescope builders and staff for their invaluable help to complete the observing run with Suprime-Cam. Data reduction/analysis in this work was in part carried out on \"sb\" computer system operated by the Astronomical Data Center (ADC) and Subaru Telescope of the National Astronomical Observatory of Japan. Use of the UH 2.2-m telescope for the observations is supported by NAOJ." }, "0801/0801.1404_arXiv.txt": { "abstract": "8-Gsps 1-bit Analog-to-Digital Converters (ADCs) were newly developed toward the realization of the wideband observation. The development of the wideband ADCs is one of the most essential developments for the radio interferometer. To evaluate its performance in interferometric observations, the time (phase) stability and frequency response were measured with a noise source and a signal generator. The results of these measurements show that the developed ADCs can achieve the jitter time less than 0.05 psec at the time interval of 1 sec, the passband frequency response with the slope less than 0.73 dB/GHz and the ripple less than 1.8 dB, and the aperture time less than 20 psec. The details of the developed ADC design, the measurement methods, and the results of these measurements are presented in this paper. ", "introduction": "A high-speed Analog-to-Digital Converter (ADC) can greatly contribute to the frontier of radio astronomy. The performance of ADC is very important because it decides the sensitivity of radio astronomical interferometric observations and the accuracy of astrometry observations and geodetic experiments. To realize the wideband observation for the Atacama Large Millimeter/Submillimeter Array (ALMA), the 2-bit ADC operating at the 4-GHz sampling rate have been developed \\citep{oki02}. A fringe spectrum that contains 20 line features as well as continuum emission at $\\sim$86.2 GHz was obtained though an interferometric observation of the Orion-KL regions using these ADCs connected to the Nobeyama Millimeter Array (NMA). This result shows that the realization of high frequency-resolution over a wide bandwidth is technically feasible \\citep{oku02}. This is the first time to obtain such a wideband astronomical fringe from a \\emph{single} correlation product of digitized signals; 20 line features can be found by a careful inspection of the spectrum. In this paper, the development of newly-developed 8-Gsps 1-bit ADC is described in Section 2, the measurement methods and the performance of the ADC in Section 3, and the conclusion in Section 4. ", "conclusions": "This paper has described the details of the developed 8-Gsps 1-bit ADCs and presented the measurement methods and the performances of the ADCs. Summary of this paper is as follows. \\begin{itemize} \\item The ADCs were developed with the commercially available DEC and DMUX at a bit rate of 10 Gbps. Input signals with power of $-20$ dBm are quantized with a sampling frequency of 8192 MHz or 4096 MHz. \\item The Allan standard deviation of the ADC boards indicates that their performance is stable in the timescale of $\\sim10^4$ sec. The jitter time in one second is estimated to be less than 0.05 psec. \\item There are slopes of $\\leq0.73$ dB/GHz in the passband responses of the developed ADC boards. The peak-to-peak values of the ripple are less than $1.8$ dB. \\item The aperture time of the developed ADCs was derived from power spectra when a strong CW signal is input into the developed ADC boards. The aperture time of the ADCs is estimated to be less than 20 psec. \\end{itemize} The performance of the developed ADCs is summarized in Table \\ref{tab:performance}." }, "0801/0801.4826_arXiv.txt": { "abstract": "It has long been known how to analytically relate the clustering properties of the collapsed structures (halos) to those of the underlying dark matter distribution for Gaussian initial conditions. Here we apply the same approach to physically motivated non-Gaussian models. The techniques we use were developed in the 1980s to deal with the clustering of peaks of non-Gaussian density fields. The description of the clustering of halos for non-Gaussian initial conditions has recently received renewed interest, motivated by the forthcoming large galaxy and cluster surveys. For inflationary-motivated non-Gaussianites, we find an analytic expression for the halo bias as a function of scale, mass and redshift, employing only the approximations of high-peaks and large separations. ", "introduction": "Constraining primordial non-Gaussianity offers a powerful test of the generation mechanism of cosmological perturbations in the early universe. While standard single-field models of slow-roll inflation lead to small departures from Gaussianity, non-standard scenarios allow for a larger level of non-Gaussianity (\\citet{BKMR04} and references therein). The standard observables to constrain non-Gaussianity are the cosmic microwave background and large-scale structure. A powerful technique is based on the abundance \\citep{MVJ00, VJKM01, Loverdeetal07, RB00, RGS00} and clustering \\citep{GW86, MLB86, LMV88} of rare events such as dark matter density peaks as they trace the tail of the underlying distribution. These theoretical predictions have been tested against numerical N-body simulations \\citep{KNS07,Grossietal07,DDHS07}. \\citet{DDHS07} showed that primordial non-Gaussianity affects the clustering of dark matter halos inducing a scale-dependent bias. This effect will be useful for constraining non-Gaussianity from future surveys which will provide a large sample of galaxy clusters over a volume comparable to the horizon size (e.g., DES, PanSTARRS, PAU, LSST, DUNE, ADEPT, SPACE, DUO) or mass-selected large clusters samples via the Sunyaev-Zel'dovich effect (e.g., ACT, SPT), considered alone or via cross-correlation techniques (e.g., ISW, lensing). Here, we resort to results and techniques developed in the 1980s \\citep{GW86, MLB86, LMV88} to extend this work and derive an accurate analytical expression for halo bias, in the presence of general non-Gaussian initial conditions, accounting for its scale, mass and redshift dependence. ", "conclusions": "We have obtained an analytic expression for the bias of dark matter halos for non-Gaussian initial conditions. The only approximations used in our approach are: {\\it i)} high peaks, {\\it i.e.} large values of $\\nu=\\delta_c(z)/\\sigma_R$ (as in the original Kaiser's formula), which essentially amounts to a limitation on the mass range over which one can apply this formula, and {\\it ii)} large separation among the halos, which is the standard assumption allowing to use linear bias. While it is true that on large scales ($k\\rightarrow 0$) the form factor, the transfer function and the window function go to unity, on the scales of interest neglecting these terms may lead to errors on $\\Delta b_h$ and therefore on $f_{\\rm NL}$ of the order of 100\\%. Comparison of these analytical findings with simulations will be presented elsewhere (Grossi {\\it et al.}, in preparation). An advantage of our approach is that it can be easily generalized to non-local and scale-dependent non-Gaussian models in which $B_{\\phi} (k_1, k_2,k_3)$ is the dominant higher-order correlation and has a general form, obtaining \\begin{eqnarray} \\frac{\\Delta b_h}{b_h}&=& \\frac{\\Delta_c(z)}{D(z)}\\frac{1}{8\\pi^2\\sigma_R^2}\\int dk_1 k_1^2 {\\cal M}_R(k_1) \\times \\nonumber \\\\ &&\\!\\int_{-1}^1\\!d\\mu {\\cal M}_R\\left(\\sqrt{\\alpha}\\right)\\frac{B_{\\phi}(k_1,\\sqrt{\\alpha},k)}{P_{\\phi}(k)}. \\label{eq: fnlk} \\end{eqnarray} Modeling the clustering of hot and cold CMB spots for non-Gaussian initial conditions, is a straightforward extension of this calculation (Heavens {\\it et al.}, in preparation). We envision that this calculation will be useful for constraining non-Gaussianity from future surveys which will provide a large sample of galaxy clusters over a volume comparable to the horizon size (e.g., DES, PanSTARRS, PAU, LSST, DUNE, ADEPT, SPACE, DUO) or mass-selected (via the Sunyaev-Zel'dovich effect) large clusters samples (e.g., ACT, SPT)." }, "0801/0801.3895_arXiv.txt": { "abstract": "\\ifdraft \\par\\hm\\par \\vspace{-13.5ex} \\fbox{\\Large\\sf Draft of \\isodate } \\par\\hm\\par \\vspace{4ex} \\fi This paper presents the sixth part to the Very Long Baseline Array (VLBA) Calibrator Survey. It contains the positions and maps of 264 sources of which 169 were not previously observed with very long baseline interferometry (VLBI). This survey, based on two 24 hour VLBA observing sessions, was focused on 1)~improving positions of 95 sources from previous VLBA Calibrator surveys that were observed either with very large a~priori position errors or were observed not long enough to get reliable positions and 2)~observing remaining new flat-spectrum sources with predicted correlated flux density in the range 100--200~mJy that were not observed in previous surveys. Source positions were derived from astrometric analysis of group delays determined at the 2.3 and 8.6~GHz frequency bands using the Calc/Solve software package. The VCS6 catalogue of source positions, plots of correlated flux density versus projected baseline length, contour plots and fits files of naturally weighted CLEAN images, as well as calibrated visibility function files are available on the Web at \\url{http://vlbi.gsfc.nasa.gov/vcs6}. ", "introduction": "\\label{s:introduction} This work is a continuation of the project of surveying the sky for bright compact radio sources. These sources can be used as phase referencing calibrators for imaging of weak objects with very long baseline interferometry (VLBI) and as targets for space navigation, monitoring the Earth's rotation, differential astrometry, and space geodesy. Precise positions of radio sources are needed for these applications. Since 1979 more than 4400 24~hour VLBI experiments have been scheduled in geodetic mode. These observations also determined parameters associated with the Earth orientation and rotation, antenna position and motions and other astrometric/geodetic parameters. By November 01, 2007, 1137 sources were observed in these experiments and 1045 of them were detected. The observations are described by \\citet{icrf98} and the latest catalog of 776 sources is given as the ICRF-Ext2 catalog \\citep{icrf-ext2-2004}. Because the astrometric and image quality of the radio sources (targets) improve with decreasing angular separation of the calibrator and the target, a higher density of sources than that from the ICRF catalogue was needed. Since 1994, 22 dedicated 24-hr experiments with the VLBA, called the VLBA calibrator survey (VCS), were made in order to increase the density of suitable VLBI phase reference sources \\citep{vcs1,vcs2,vcs3,vcs4,vcs5}. In these experiments, 3601 sources with declinations $> -40^\\circ$ were observed, including 764 objects previously observed under geodetic programs, and 3301 of them were detected. The catalogue of source positions derived from a single least square (LSQ) solution using all geodetic and VCS observations forms the pool of sources with positions at a milliarcsec level of accuracy that is widely used for phase referencing observations, for statistical analysis, and for other applications. Improving the precision of this catalogue and an increase in the number of objects is an important task. In this paper we present an extension of the VCS catalogue, called the VCS6 catalogue. The objectives of this campaign were to improve the catalogue of compact radio sources. Our approach to this problem was 1)~to improve positions of sources observed in the previous VCS campaigns that had a)~poor a~priori positions; b)~were not observed long enough in the first two VCS1 24-hour experiments recorded at the 64~Mbit/sec rate; and 2)~to get positions of new sources. The new sources were taken from the lists of a)~intra-day variable sources observed in the framework of the MASIV survey \\citep{masiv-2003}, not observed before with the VCS, and b)~leftovers from the list of candidate sources prepared for the VCS5 campaign \\citep{vcs5} that were not observed due to lack of resources. Since the observations, calibrations, astrometric solutions and imaging are similar to that of VCS1--5, most of the details are described by \\cite{vcs1}, \\cite{vcs3}, and \\cite{vcs4}. In \\S\\ref{s:selection} we describe the strategy for selecting 285 candidate sources observed in two 24~hour sessions with the VLBA. In \\S\\ref{s:obs_anal} we briefly outline the observations and data processing. We present the VCS6 catalogue in \\S\\ref{s:catalogue}, and summarize our results in \\S\\ref{s:summary}. ", "conclusions": "\\label{s:summary} The VCS6 Survey has added 169 new compact radio sources not previously observed with VLBI and significantly improved coordinates of 95 other objects. Among the new sources, 103 objects turned out to be suitable as phase referencing calibrators and as target sources for geodetic applications. Their coordinates have position accuracy better than 5~mas." }, "0801/0801.1068_arXiv.txt": { "abstract": "We present CCD photometric and mass function study of 9 young Large Magellanic Cloud star clusters namely \\CLN. The \\ubvri data reaching down to $V \\sim$ 21 mag, are collected from 3.5-meter NTT/EFOSC2 in sub-arcsec seeing conditions. For NGC 1767, NGC 1994, NGC 2002, NGC 2003, NGC 2011 and NGC 2136, broad band photometric CCD data are presented for the first time. Seven of the 9 clusters have ages between 16 to 25 Myr while remaining two clusters have ages $32\\pm4$ Myr (NGC 2098) and $90\\pm10$ Myr (NGC 2136). For 7 younger clusters, the age estimates based on a recent model and the integrated spectra are found to be systematically lower ($\\sim$ 10 Myr) from the present estimate. In the mass range of $\\sim 2 - 12$ $M_{\\odot}$, the MF slopes for 8 out of nine clusters were found to be similar with the value of $\\gamma$ ranging from $-1.90\\pm0.16$ to $-2.28\\pm0.21$. For NGC 1767 it is flatter with $\\gamma = -1.23\\pm0.27$. Mass segregation effects are observed for NGC 2002, NGC 2006, NGC 2136 and NGC 2098. This is consistent with the findings of \\citet{kontizas98} for NGC 2098. Presence of mass segregation in these clusters could be an imprint of star formation process as their ages are significantly smaller than their dynamical evolution time. Mean MF slope of $\\gamma = -2.22\\pm0.16$ derived for a sample of 25 young ($\\le 100$ Myr) dynamically unevolved LMC stellar systems provide support for the universality of IMF in the intermediate mass range $\\sim 2-12\\ M_{\\odot}$. ", "introduction": "\\label{sec:intro} The distribution of stellar masses that form in one star-formation event in a given volume of space is called initial mass function (IMF). Some theoretical studies consider that the IMF should vary with the pressure and temperature of the star-forming cloud in such a way that higher-temperature regions ought to produce higher average stellar masses while others have exactly opposite views (see \\citealt{larson98, elmegreen00} and references therein). Detailed knowledge of the IMF shape in different star forming environments is therefore essential. One would like to know whether it is universal in time and space or not. For this, in a galaxy, its young (age $\\le$ 100 Myr) star clusters of different ages, abundance etc. need to be observed, as they contain dynamically unevolved, (almost) coeval sets of stars at the same distance with the same metallicity. For a number of such reasons, populous young star clusters of the Large Magellanic Clouds (LMC) are the most suitable objects for investigating the IMF. They provide physical conditions not present in our Galaxy e.g. stellar richness, metallicity and mass ranges (see \\citealt{sagar93, sagar95} and references therein). Unlike the galactic counterparts, where corrections for interstellar absorption are not always trivial since it could be large as well as variable \\citep{sagar87, yadav01, kumar04}, for LMC star clusters it is relatively small. Its treatment is therefore not a problem. Furthermore, choosing young (age $\\le 100$ Myr) clusters reduces the effects of dynamical evolution on their MF. The present day mass function of these stellar systems can therefore be considered as the IMF. The study of young LMC star clusters is thus important for providing the answer to the question of universality of the IMF. Both ground and Hubble Space Telescope (HST) observations have therefore been obtained (see \\citealt{sagar91a, brocato01, matteucci02} and references therein) for a few of the large number of young LMC star clusters \\citep{bica99}. The potential offered by them has not been fully utilised as still a large fraction of them are unobserved. \\begin{figure} \\centering \\includegraphics[width=9cm]{radec.eps} \\caption{Small dots show the location of identified LMC star clusters from the catalog of \\citet{bica99}. A sky area of about 8\\degr $\\times$ 8\\degr is shown centered around the optical center ($\\alpha_{\\rm J2000} =5^h 20^{m} 56^{s}$, $\\delta_{\\rm J2000} = -69^{\\degr} 28^{\\arcmin} 41^{\\arcsec}$) of the LMC. The bar region is clearly seen. The target clusters are shown with filled triangles.} \\label{fig:radec} \\end{figure} In this paper we derive mass function (MF) slopes using new broad band $BVRI$ CCD photometric observations of the stars in 9 young LMC star clusters namely \\CLN. Their integrated photometric colours indicate that all of them belong to SWB \\citep{searle80} class of 0 or I and hence are very young with ages $\\le 30$ Myr \\citep{elson95, bica96} except NGC 2136 which belongs to SWB class of III indicating and age between 70\\,-\\,200 Myr. Table \\ref{tab:sample} lists the relevant information available prior to this study. All the clusters are rich indicating higher stellar density \\citep{bica95} and thus most suitable for the MF study. Except NGC 2011, all are elliptical in size with major axes diameters ranging from 1\\farcm3 to 2\\farcm8. Except for NGC 2002 and NGC 2098, all others are candidate members of either a pair or a multiple system (see \\citealt{dieball02}). Locations of the target clusters are shown in Fig. \\ref{fig:radec}. Most of them lie towards north-east side of the LMC bar harboring young star forming regions in contrast to the intermediate age (1 - 3 Gyr) cluster field of the bar. NGC 1767 lies south-west region of the LMC bar. Being spread over a wide region ($\\sim$ 5\\degr $\\times$ 10\\degr), the sample may reflect different star forming environments. It is therefore suitable for testing the universality of the IMF. When CCD observations were carried out in 1990, no detailed photometric observations and MF studies had been published. However, in the mean time some CCD photometric observations have been published for a few of the clusters under study. A brief description of the previous work on the clusters under study is given below. \\input{./sample.tab} \\subsection{Previous work} \\begin{itemize} \\item {\\bf NGC 1767.} This, a member of triple star cluster system, is located in the OB association LH 8. Integrated ($U-B$) and ($B-V$) colors indicate that the cluster is young with an age of $\\sim$ 10 Myr. \\item {\\bf NGC 1994.} This, located in LMC DEM 210 region, is a member of a 5-cluster system. Its irregular size is largest amongst them. An age of about 5 - 30 Myr has been derived for the cluster from its integrated photometric colour observations. \\item {\\bf NGC 2002.} This single cluster is located in the OB association LH 77 in the supergiant shell LMC 4 region. The cluster center is condensed, but the outer part is resolved. Integrated light observations indicate an age of $\\sim$ 10 - 30 Myr along with the presence of a few red supergiants \\citep{bica96}. \\item {\\bf NGC 2003.} Integrated photometric observations indicate an age of 10 - 30 Myr for this cluster located in the Shapley III region of the LMC. Its shape on the photographic image is elongated with resolved outer parts. \\item {\\bf NGC 2006 and SL 538.} This binary star cluster is located in the northwestern part of the OB association LH 77 in supergiant shell LMC 4. The clusters are separated by $\\sim 55''$ on the sky corresponding to a linear separation of 13.3 pc at the distance of LMC. Integrated photometric observations obtained by \\citet{bhatia92} and \\citet{bica96} indicate a similar age for both the clusters. Using low-resolution objective prism spectra and integrated IUE spectra, \\citet{kontizas98} suggested that this binary cluster may merge in $\\sim$ 10 Myr. Broad band and H$_{\\alpha}$ CCD photometric observations were obtained by \\citet{dieball98}. Based on the colour-magnitude diagrams (CMDs) of the clusters they derived an age of $18\\pm2$ Myr for SL 538 and of $22.5\\pm2.5$ Myr for NGC 2006. The MF slopes obtained for both the clusters were consistent with that of \\citet{salpeter55} and indicated similar total masses. These studies thus indicates near-simultaneous formation of the cluster pair in the same giant molecular cloud. \\item {\\bf NGC 2011.} This is located in the OB association region LH 75. Its age estimated from the integrated photometric observations is between 10 to few tens of Myr. Its photographic image indicates that it is elongated, fairly condensed and partly resolved cluster. A recent analysis of its stellar content using HST observations reveal that it has two parallel main sequence branches, and may be a binary system \\citep{gouliermis06}. However, the analysis also indicate that both populations might have formed in a single star forming event as the redder stars are situated in the central half arcmin region and are thought to be embedded in the dust and gas, while the blue stars are spread in the outer region up to 1 arcmin. \\item {\\bf NGC 2098.} This is another single cluster amongst the objects under study. The first $BR$ broad band CCD photometric observations have been presented by \\citet{kontizas98}. They derived an age of 63 - 79 Myr and found strong evidence for mass segregation in agreement with their earlier studies based on the photographic observations. However, poor quality of their CCD data was indicated by the authors. \\item {\\bf NGC 2136.} This is the brighter component of the young binary globular cluster NGC 2136/ NGC 2137 in the LMC. The angular separation between the components is about 1\\farcm3. \\citet{hilker95} using Stromgren CCD photometry of the clusters indicates their common origin. They indicate an age of 80 Myr and metallicity [Fe/H] = $-0.55\\pm0.06$ dex for the cluster while \\citet{dirsch00} derive an age of $100\\pm20$ Myr but the same metallicity. The cluster contains a number of Cepheids as well as red giants. \\end{itemize} The present CCD observations, in combination with earlier observations, have been used to estimate and/or interpret the interstellar reddening to the cluster regions, ages and mass functions of the clusters. Section 2 deals with the observational data, reduction procedures and comparisons with the published photometric data. In section 3, we analyse the stellar surface density profiles, CMDs and MFs of the sample clusters. Last section is presented with the results and discussions. \\input{./obslog.tab} ", "conclusions": "We present \\ubvri CCD data obtained from 3.5-meter ESO NTT/EFOSC2 observations for 9 young Large Magellanic Cloud star clusters namely \\CLN\\ and their nearby field regions reaching down to $V \\sim 20$ mag for $\\sim$ 6400 stars altogether. They are the first accurate broad band CCD photometric data for all the clusters except for the binary cluster NGC 2006 and SL 538. The observations are made in a region of $\\sim 2\\arcmin \\times 2\\arcmin$ around the cluster center. The data were collected during Jan 10 to Jan 13, 1990 in good seeing conditions ranging from 0\\farcs7 to 1\\farcs0 and reduced using DAOPHOT and MIDAS softwares. Photometric calibrations are done using Landolt (1992) stars and the zero point accuracy is better than $0.02$ mag. Photometric errors become large (\\ga 0.1 mag) for stars fainter than $V = 20$ mag. We examine radial density profiles, general features of the main sequence and estimate age and reddening for individual clusters using Padova isochrones. The various CMDs of the clusters under study were used to estimate their MF, age and reddening. In order to study radial variation in MF, the LFs are derived for inner, outer, and entire cluster regions. Due to compactness of the clusters, such study could not be carried out for the core regions of the clusters. The LFs are corrected for both data incompleteness and field contamination. The main conclusions of the present study are as follows. \\begin{enumerate} \\item Seven of the nine clusters have ages $\\le 25$ Myr, while the remaining two clusters have ages of $32\\pm4$ Myr (NGC 2098) and $90\\pm10$ Myr (NGC 2136). Our age estimates for NGC 2006 and SL 538 were found to be consistent with the previous $BVR$ photometric estimate by \\citet{dieball98}. For NGC 2098, our estimates are lower by about 30 Myr than \\citet{kontizas98}. Thus, the ages of all the clusters in our sample are significantly lower than their typical dynamical ages of a few 100 Myrs. \\item For younger ($\\le$ 25 Myr) clusters, the age estimates based on a recent population synthesis models by \\citet{wolf07} and integrated spectra are systematically lower by about 10 Myr than the present age estimates based on CMDs. \\item Assuming an LMC distance modulus of 18.5 mag, the derived reddening for the clusters in our sample was found to be consistent with that derived from HI emission and 100 $\\mu m$ all sky dust maps. \\item In the mass range of $2 - 12\\ M_{\\odot}$, the MF slopes for 8 out of 9 sample clusters were found to be similar with values of $\\gamma$ ranging from $-1.90\\pm0.16$ to $-2.28\\pm0.21$. For NGC 1767 the slope was found to be significantly shallower with $\\gamma = -1.23\\pm0.25$. The present MF values are consistent with those derived by \\citet{dieball98} for NGC 2006 and SL 538. \\citet{selman05} studied the star formation history and IMF of the field population of 30 Doradus super association and found that it has a Salpeter slope in the mass range of 7 to 40 $M_{\\odot}$. \\item Mass function slopes of the inner and outer cluster regions indicates the presence of mass segregations in NGC 2002, NGC 2006, NGC 2136 and NGC 2098. For NGC 2098, \\citet{kontizas98} derive the dynamical relaxation time, $T_{\\rm e}$ between 640 to 1050 Myr. This may indicate that the value of $T_{\\rm e}$ for LMC star clusters could be few hundreds of Myr. The ages of LMC star clusters under study are therefore significantly smaller than their dynamical relaxation time. Consequently, observed mass segregation in these clusters is probably primordial in nature. A compilation of both ground and space based observations of extremely young galactic and MC star clusters (cf. \\citealt{hunter95}; \\citealt{sagar88}; \\citealt{hillenbrand98}; \\citealt{chen07} and references therein) indicates presence of mass segregation in most of them, although to varying degrees. All these indicate that in most of the young star clusters located in different galaxies, mass segregation effects are observed and most likely they are imprint of star formation processes. \\item A mean MF slope of $\\gamma = -2.22\\pm0.16$ derived for a sample of 25 young ($< 100$ Myr) stellar systems in LMC provide support for the universality of IMF in the intermediate mass range $\\sim 2-10\\ M_{\\odot}$. An IMF study of the 30 Doradus star forming region of LMC by \\citet{selman05} also support this conclusion. \\end{enumerate}" }, "0801/0801.0402_arXiv.txt": { "abstract": "We present the Spitzer Infrared Spectrograph (IRS) spectrum of SR20, a 5--10 AU binary T Tauri system in the $\\rho$ Ophiuchi star forming region. The spectrum has features consistent with the presence of a disk; however, the continuum slope is steeper than the $\\lambda^{-4/3}$ slope of an infinite geometrically thin, optically thick disk, indicating that the disk is outwardly truncated. Comparison with photometry from the literature shows a large increase in the mid-infrared flux from 1993 to 1996. We model the spectral energy distribution and IRS spectrum with a wall $+$ optically thick irradiated disk, yielding an outer radius of 0.39$_{+0.03}^{-0.01}$ AU, much smaller than predicted by models of binary orbits. Using a two temperature $\\chi^2$ minimization model to fit the dust composition of the IRS spectrum, we find the disk has experienced significant grain growth: its spectrum is well-fit using opacities of grains larger than 1 $\\mu$m. We conclude that the system experienced a significant gravitational perturbation in the 1990s. ", "introduction": "Classical T Tauri stars are typically surrounded by optically thick accretion disks, indicated by strong and broad H$\\alpha$ emission lines and by a characteristic infrared excess. Over the last twenty years, observations have revealed that many of these young stars are components in binary, and sometimes higher order, systems \\citep[e.g.][]{simon92, ghez93}. A binary system may produce up to three distinct disks: a circumprimary, circumsecondary, and circumbinary disk \\citep[][(AL94; JM97)]{artymowicz94, jensen97}. Gravitational interactions between binaries affect the structure of their disks; in particular, the maximum outer radius for a circumstellar disk in a binary system is approximately 0.18 to 0.4 times the semi-major axis (AL94), depending on the mass ratio and orbital eccentricity. SR20 is one such close binary in the $\\rho$ Ophiuchi star forming region. Its components have an angular separation ranging from 0.038 to 0.071\\arcsec \\citep{ghez93, ghez95} which, at a distance to Ophiuchus of 140 pc \\citep{dz99} corresponds to a projected linear separation of 5.3--9.9 AU. The secondary is 2.2 magnitudes fainter than the primary at 2.2 $\\mu$m \\citep{ghez93} and therefore is less massive than the primary, diskless, or both. H$\\alpha$ emission has been detected from SR20 with equivalent widths ranging from 15 to 21 \\AA \\citep[][(W05)]{r80, ba92, wilking05}. Further observations by \\citet{tbc03} did not find a spectro-astrometric signature in the H$\\alpha$ line, indicating that there is little to no accretion in the secondary relative to the primary. Based on the lack of evidence for a circumsecondary disk, we assume that the infrared excess of SR20 is dominated by a circumprimary accretion disk. Here we present Spitzer Space Telescope Infrared Spectrograph (IRS) \\citep{houck04} observations of the SR20 system. We model the structure of the system and the dust it contains and analyze the nature of the SR20 system. While instances of sculpture of proto-planetary accretion disks by planetary or stellar companions \\emph{within} the disk are becoming widely known \\citep[e.g. JM97,][]{furlan07}, SR20 is an apparently rarer example of sculpture from \\emph{without}. ", "conclusions": "\\label{discussion} The SED, system geometry, and dust composition shed light on the nature of the SR20 system, while simultaneously raising further questions. The IRS spectrum is well-fit by our composite wall $+$ circumprimary disk truncated beyond 0.39 AU. Outward truncation could be caused by the orbit of a companion or lack of replenishment from a circumbinary disk. Using the tidal truncation models of AL94, the range of likely eccentricities (0.2-0.3) and semi-major axes (8-12 AU) of the A-B orbit are not extreme enough to cause truncation. Even with the lower limit on the semi-major axis, 5 AU, and the highest order resonances and eccentricities tested in AL94, the minimum stable outer truncation radius caused by the known binary is 0.9 AU, considerably larger than the 0.39 AU we derive (AL94). However, the AL94 models may not be appropriate for a disk with such a small ratio between the outer and inner radii, so we cannot {\\it absolutely} rule out truncation by the known binary, at least on this basis. Alternatively, it may be that SR20 is a hierarchical triple. A companion object with an eccentricity $<$ 0.8 could orbit in the region between 0.98 AU and 2.17 AU from the primary, causing a disk truncation at 0.39 AU. Such a close, low mass companion would be difficult to detect. Based on the 0.39 AU truncation radius, we speculate that replenishment of the circumprimary disk by a circumbinary disk occurred in the past. A disk between 0.1 and 10 AU would accrete on a timescale of 1,000--100,000 years \\citep{quillen04} and the median age of the Ophiuchus cluster is 2.1 Myr (W05). SR20 has been observed at 800, 850, and 1300 $\\mu$m with only 3 $\\sigma$ upper limit results, implying less than $5.4$ $\\times$ $10^{4}$ $M_{lunar}$ (540 $M_{lunar}$ of dust) in disk material \\citep{aw07}. Low circumbinary disk masses are typical of binaries with semi-major axes between 1 and 50 AU \\citep{jensen94}, so submillimeter non-detection does not indicate a depleted circumbinary disk; however, it is unlikely that lack of replenishment is responsible for the 0.39 AU truncation radius. A nearby companion truncating the circumprimary disk is perhaps also the best explanation for the photometric variability. Before 1993, the infrared excess in SR20 started in the N band, indicating that there were few small dust grains close to the central star. However, the H$\\alpha$ emission indicates that a gaseous accretion disk, perhaps mixed with much larger planetesimals, was present. Between 1993 and 1996, the infrared excess increased sharply, indicating an increase in the amount of dust in the disk, which could be explained by a gravitational interaction between the known binary and a third, unseen, body, causing the unseen companion to orbit closer to the primary, perturbing the circumprimary disk. Coupling between the gas and dust could confine the newly created dust to the accretion disk. The appeareance of the observed excess over 3 years, preceded by 20 years without an excess and followed by 10 years of little variation in the observed excess, seems consistent with truncation by a nearby source with a shorter period than is plausible for the known companion. The silicate dust composition of the system is not typical of a Class II object; our fit indicates that the optically thin portion of the disk is comprised primarily of large grains ($>$ 1 $\\mu$m). Although having a large fraction of large grains is consistent with debris disks or advanced dust processing in the inner regions of a protoplanetary disk \\citep{chen06}, this is the first time we have seen a Class II disk with negligible submicron grains. Taken with the aforementioned explanations of the 1993--1996 variability, the large grains could support a collision scenario: gravitional interactions between the known secondary and the unseen companion cause collisions between planetesimals, producing dust grains that are detected as the current disk. If there were a population of planetesimals -- invisible in our spectra -- comprising an extension of the disk beyond the outer truncation radius of the dust, truncation by the known companion would be more probable. Near infrared interferometry and orbital monitoring would be necessary to determine the nature of this complex system more precisely." }, "0801/0801.2631_arXiv.txt": { "abstract": "We present theoretical evolutionary sequences of intermediate mass stars (M=3-6.5$\\msun$) with metallicity Z=0.004. Our goal is to test whether the self-enrichment scenario by massive Asymptotic Giant Branch stars may work for the high metallicity Globular Clusters, after previous works by the same group showed that the theoretical yields by this class of objects can reproduce the observed trends among the abundances of some elements, namely the O-Al and O-Na anticorrelations, at intermediate metallicities, i.e [Fe/H]=-1.3. We find that the increase in the metallicity favours only a modest decrease of the luminosity and the temperature at the bottom of the envelope for the same core mass, and also the efficiency of the third dredge-up is scarcely altered. On the contrary, differences are found in the yields, due to the different impact that processes with the same efficiency have on the overall abundance of envelopes with different metallicities. We expect the same qualitative patterns as in the intermediate metallicity case, but the slopes of some of the relationships among the abundances of some elements are different. We compare the sodium--oxygen anticorrelation for clusters of intermediate metallicity (Z$\\approx 10^{-3}$) and clusters of metallicity large as in these new models. Although the observational data are still too scarce, the models are consistent with the observed trends, provided that only stars of M$\\simgt$5\\msun\\ contribute to self--enrichment. ", "introduction": "Spectroscopic investigations of Globular Clusters (GC) show star to star differences in their surface chemistries (Kraft 1994). These inhomogeneities trace clear abundance patterns involving all the light elements: sodium is correlated to aluminium and anticorrelated to oxygen, whereas magnesium is anticorrelated with aluminium (Carretta 2006), though in some clusters the existence of this latter relationship is still under debate (Cohen \\& Mel\\'endez 2005). A common result of these analysis is the approximate constancy of the overall CNO abundances (Ivans et al. 1999). The detection of such anomalies in scarcely evolved stars, like Turn-Off (TO) and Sub-Giant Branch (SGB) stars (Gratton et al 2001), ruled out the possibility of any in situ mechanism as a possible unique explanation (Denissenkov \\& Weiss 2004), and pointed in favour of a self-enrichment scenario, i.e. that the stars with the anomalous chemistry formed from the ashes of a previous stellar generation, which polluted the interstellar medium with matter which had been processed via the CNO cycle. Ventura et al. (2001) indicated the massive Asymptotic Giant Branch (AGB) stars as likely candidates, because with appropriate hypothesis concerning convection modelling they can easily achieve Hot Bottom Burning (HBB), i.e. the bottom of their envelope becomes so hot to ignite strong nucleosynthesis, whose products, convected to the surface of the star, are ejected into the interstellar medium via the strong winds suffered by these sources (Ventura \\& D'Antona 2005); also, strong HBB favours high mass loss, reduces the number of thermal pulses (TP) experienced by the stars during their life, thus diminishing the number of third dregde-up (TDU) episodes, and keeping approximately constant the C+N+O mass fraction in the envelope, for a reasonably large range of initial masses. This conclusion is at odds with the results by Fenner et al. (2004), who, based on AGB models calculated with the traditional, low efficiency, mixing length theory (MLT, Vitense 1953) convective model, found that oxygen can be hardly depleted at the surface of these stars, and the CNO sum is expected to largely exceed unity. Recently, an alternative self-enrichment scenario for GCs was proposed by Maeder \\& Meynet (2006), Prantzos \\& Charbonnel (2006), and described in details by Decressin et al. (2007): in this case the enrichment of the interstellar medium is produced by the envelopes of fast rotating massive stars. In order to choose between the AGB and the massive stars self-enrichment, it is necessary to explore in detail both models. In particular, the variation of predictions of yields with metal abundance must be examined. Ventura \\& D'Antona (2008, hereinafter VD08) showed that the enrichment by AGBs can work in the case of intermediate metallicity GCs, since the ejecta of their theoretical models, calculated with a metallicity Z=0.001, with initial masses in the range 5-6$\\msun$, are in qualitative and quantitative agreement with the abundance patterns observed in scarcely evolved stars of M3, M13, M5, NGC6218, NGC6752. The goal of the present paper is to investigate whether the self-enrichment scenario hypothesis may work also in the case of metal rich GCs, having metallicities [Fe/H]$\\sim -0.7$. We present a new set of AGB models with metallicity Z=4$\\times 10^{-3}$, typical of metal rich GCs like 47Tuc and M71. We discuss the possibility of reproducing the O-Al and O-Na anticorrelations, and determine the expected slope of these relationships compared to the intermediate metallicity case. \\begin{figure} \\includegraphics[width=.48\\textwidth]{fig1.eps} \\caption{The variation of the temperature at the bottom of the convective envelope as a function of the core mass at the beginning of the AGB phase (Top) and of the initial mass (Bottom) of AGB models with two different metallicities. The temperatures refer to the phase of maximum luminosity during the AGB evolution.} \\label{phys} \\end{figure} ", "conclusions": "We present new AGB models with metallicity Z=0.004, to test the hypothesis of AGB self-enrichment in high metallicity GCs and compare the results with previous models having Z=0.001 (VD08). We find that the main physical features of the AGB evolution of the intermediate mass stars do not change by increasing the metallicity, the main effects being only a shift towards higher masses of the mass-luminosity relationship; models with the same core mass and different Z follow very similar evolution, in terms of the temperature reached at the bottom of their convective envelope. The theoretical yields of the most massive AGB models are in agreement with the aluminum, sodium and oxygen abundances of the most anomalous stars (those showing the strongest depletion of oxygen ---and excluding the giant in which the very low O should be attributed to deep extramixing) in the few high metallicity GCs so far examined; a reasonable dilution scheme gives results consistent with the O--Na and C--N anticorrelations. Yet, other confirmation is needed, since the data currently available involve only few stars per cluster. One robust prediction of this investigation, which is independent of all the uncertainties associated to the proton capture cross sections by Ne-Na-Al nuclei, is that with increasing metallicity we expect a higher slope of the O-Na anticorrelation, and a similar slope (though less extended) of the O-Al trend." }, "0801/0801.0634_arXiv.txt": { "abstract": "The size distribution of mini-filaments in voids has been derived from the Millennium Run halo catalogs at redshifts $z=0,0.5,1$ and $2$. It is assumed that the primordial tidal field originated the presence of filamentary substructures in voids and that the void filaments have evolved only little, keeping the initial memory of the primordial tidal field. Applying the filament-finding algorithm based on the minimal spanning tree (MST) technique to the Millennium voids, we identify the mini-filaments running through voids and measure their sizes at each redshift. Then, we calculate the comoving number density of void filaments as a function of their sizes in the logarithmic interval and determine an analytic fitting function for it. It is found that the size distribution of void mini-filaments in the logarithmic interval, $dN/d\\log S$, has an almost universal shape, insensitive to the redshift: In the short-size section it is well approximated as a power-law, $dN/d\\log S \\approx S$, while in the long-size section it decreases exponentially as $dN/d\\log S \\approx \\exp(-S^{\\alpha})$. We expect that the universal size distribution of void filaments may provide a useful cosmological probe without resorting to the rms density fluctuations. ", "introduction": "\\label{intro} In the classical theory of structure formation, it was generally believed that the gravity is fully responsible for the formation and evolution of the large scale structure in the universe. Recently, however, it has been realized that the overall characteristics of the large scale structures cannot be understood only in terms of the gravitational influence. It has been pointed out that the tidal field provides a driving force in establishing the observed large scale filamentary structures in the universe \\citep{bon-etal96,lee-evr07, par-lee07b,hah-etal07,ara-etal07}. The cosmic voids are most vulnerable to the tidal influence from the surrounding matter distribution due to their extreme low-density \\citep{sah-sha96,sha-etal04,sha-etal06,lee-par06,par-lee07b}. The tidal squeezing and distortion effect tends to deviate the void shapes from spherical symmetry and could sometimes lead even to the collapse and disappearance of the voids \\citep{sah-etal94,sah-sha96,sha-etal04}. The presence of the filamentary structures in voids marks the most striking evidence for the strong tidal influence on voids. The anisotropy in the spatial distribution of void halos is induced by their alignments with the principal axes of the tidal tensors \\citep{pee01}. Since the void halos evolve very little and most particles remain primordial in voids \\citep{ein06}, the void mini-filaments are pristine, keeping well the initial memory of the spatial coherence of the primordial tidal effect, unlike the large scale filaments which also originated from the large-scale coherence of the primordial tidal field but have undergone highly nonlinear processes in the subsequent evolution. Hence, the void filaments are the most idealistic probe of the primordial tidal field and its effect on voids. The size distribution of void filaments should be a function of the strength of the tidal effect and the host void sizes. The maximum size of a void filament cannot exceed the size of its host void. But, even when a host void has a large size, it would not have long-size filaments if there is no tidal influence. Here, we attempt to derive the size distributions of the void filaments at various redshifts and to explore their statistical properties. Before presenting our results, however, we would like to caution the readers for the difficulty in using the void filaments as a probe of the primordial tidal field. Unlike the galaxy clusters, there is no unique way to define voids and their filaments. Recently, the seminal work of \\citet{col-etal08} compared thoroughly various void-finding algorithms and showed clearly that different void finders different void characteristics, under-density profiles, void galaxies etc. Thus, our results are likely to be dependent on our specific choice of the void-finder. The organization of this paper is as follows. In \\S \\ref{void}, we briefly describe the void catalog was obtained from the Millennium Run simulations. In \\S \\ref{filament}, we explain how the void filaments are identified from the void catalog. In \\S \\ref{result}, we derive numerically the size distributions of void filaments and determine an analytic fitting formula for it. In \\S \\ref{end}, we discuss the final results and assess the future work. ", "conclusions": "\\label{end} It is generally thought that the large scale structure of the universe provides a window on the initial condition of the early universe. In previous cosmological studies, it was the halo mass function that has been most highlighted as a statistical tool for probing cosmology with the large scale structure \\citep{pre-sch74,she-tor99,she-etal01,jen-etal01,ree-etal03}. The most advantageous aspect of the halo mass function is that it can be written in a simple universal form, independent of redshift \\citep{she-tor99}. Yet, the halo mass function has a generic weakness as a cosmological probe: its dependence on the cosmological parameters comes indirectly from the dependence of the halo mass on the rms fluctuation, $\\sigma(R)$, of the linear density field smoothed on a scale radius of $R$. It is desirable to develop another statistical tool which can overcome the weakness of the halo mass function. Recently, several authors have noted the possibility of probing cosmology with cosmic voids \\citep[e.g.,][]{she-van04,par-lee07a,pla-etal08}. Yet, no new statistical tool that can really compete with the halo mass function has been suggested so far. For instance, in our previous work \\citep{par-lee07a}, we have proposed that the cosmic voids provide another window on cosmology. Given that the void ellipticities are induced by the primordial tidal effect which depends on the background cosmology, we have shown analytically that the void ellipticity distribution can be used to constrain the cosmological parameters. However, it has turned out that the void ellipticity distribution suffers from the same weakness that the halo mass function has. Its dependence on cosmology is secondary, resulting from the dependence of the void scale on the linear rms density fluctuation, $\\sigma(R)$. Furthermore, the void ellipticity distribution cannot be written in a universal form unlike the halo mass function. Therefore, true as it is that the void ellipticity distribution provides another way to determine the cosmological parameters with the large scale structure, it is not competitively compared with the halo mass function. In this work, we have for the first time quantified the effect of the tidal field on cosmic voids by the sizes of the void filaments. It is shown that the size distribution of void filaments is almost independent of redshift, having a simple universal form. A crucial implication of our result is that the number density of long filaments in voids should depend very sensitively on the background cosmology since the sizes of the void filaments reflect the primordial tidal field. Our result that the size distribution of void filaments is almost independent of redshift in spite of the well known fact that the sizes of voids grow very strongly with redshift can be understood as follows. It is true that the comoving sizes of voids keep growing as the voids expand faster than the rest of universe due to their extreme low density. However, the comoving sizes of void filaments should not necessarily keep growing with redshift. But for the effect of the tidal field, the spatial distribution of void halos would be isotropic due to the gravitational rarefaction effect caused by the fast expansion of voids. In other words, while the gravitational rarefaction effect tends to make the spatial distribution of void halos isotropic, the tidal effect on voids tends to increase the anisotropy in the distribution of void halos, resulting in growth of void filaments. Therefore, as far as the tidal effect exists and counteracts the expansion effect, the sizes of void filaments will also grow as the sizes of their host voids grow with redshift. However, when the void size grows large enough to finally reach its threshold at which the expansion effect overcomes the tidal effect, the degree of anisotropy in the spatial distribution of void halos will diminish and accordingly the sizes of void filaments will stop growing. Since this critical threshold of void size is determined by the single condition that the expansion effect overcomes the tidal effect, its value should be redshift-independent. At high-redshift when the sizes of voids were small, the strong dominant tidal effect increases the anisotropy in the spatial distribution of void halos, increasing the sizes of void filaments. At low-redshift when the sizes of voids are large, the strong expansion effect prevents the sizes of void filaments from keep growing. Hence, the universal size distribution of void filaments reflects the fact that the sizes of void filaments are regulated by the counter-balance between the expansion effect and the tidal effect. The fitting formula, eqs. (\\ref{eqn:fit1})-(\\ref{eqn:fit2}), indicate that the number density of void filaments, $dN/d\\log\\nu$, behaves like a power-law with power index of $n=1$ in the short filament section while it decreases exponentially in the long filament section. A crucial implication of our result is that the exponential decrease of the abundance of long-size filaments in voids should be very sensitive to the key cosmological parameters. Especially it is expected to depend sensitively on the amplitude of the linear power spectrum, $\\sigma_{8}$. In our previous work, we have already shown that the void ellipticity caused by the tidal effect increases with the value of $\\sigma_{8}$ \\citep{par-lee07a}. Accordingly, we expect that the number density of long filaments in voids would increase as the value of $\\sigma_{8}$ increases. Unlike the void ellipticity distribution whose dependence on $\\sigma_{8}$ is weak and indirect through its dependence on the rms density fluctuation, equation (\\ref{eqn:fit1}) suggests that the size distribution of void filaments should depend strongly on the power spectrum amplitude. To use the size distribution of void filaments as a cosmological probe, however, there should be some additional tasks to be done in the future. First, the mass-to-light bias and the redshift distortion effects have to be account for. What one can observe is not halos in real space but galaxies in redshift space. The size distribution of void filaments measured from the galaxy catalogs in redshift space could differ from the current result. Thus, it will be very necessary to investigate how the bias and the redshift distortion effect change the size distribution of void filaments. It is expected that the sizes of void filaments may increase in redshift space since the redshift space distortion effect tend to increase the anisotropy in the spatial distribution of void galaxies. Meanwhile, the matter-to-light bias might decrease the sizes of void filaments, compensating for the redshift distortion effect, since the massive halos are found to be distributed less anisotropically (Park \\& Lee 2009 in preparation). Henceforth, the overall size distribution of void galaxy filaments in redshift space might be similar to that in real space, due to the competition between the two effects (in private communication with van de Weygaert). Second, a general analytic formula for the size distribution of void filaments in an arbitrary cosmology is desirable to derive from physical principles. If an analytic model is found, it would allow us to understand the true physical meaning of eqs. (\\ref{eqn:fit1})-(\\ref{eqn:fit2}). Third, it should be worth examining whether the results depend on the choice of the filament-finding and the void-finding algorithms. As mentioned in \\S 1, it has been found that different void-finders provide different void properties and the HV02 algorithm that we used here has a tendency to find larger voids than most of the other void-finders \\citep{col-etal08}. Furthermore, our results are not general ones but valid only for the specific cosmology of the Millennium Run simulation. Therefore, it is required to test the universality of the size distribution of void filaments with other void-finders and with other cosmologies as well. We plan to address these issues and report the results elsewhere in the near future." }, "0801/0801.4420_arXiv.txt": { "abstract": "In cyclic cosmology based on phantom dark energy the requirement that our universe satisfy a CBE-condition ({\\it Comes Back Empty}) imposes a lower bound on the number $N_{\\rm cp}$ of causal patches which separate just prior to turnaround. This bound depends on the dark energy equation of state $w = p/\\rho = -1 - \\phi$ with $\\phi > 0$. More accurate measurement of $\\phi$ will constrain $N_{\\rm cp}$. The critical density $\\rho_c$ in the model has a lower bound $\\rho_c \\ge (10^9 {\\rm GeV})^4$ or $\\rho_c \\ge (10^{18} {\\rm GeV})^4$ when the smallest bound state has size $10^{-15}$m, or $10^{-35}$m, respectively. ", "introduction": "\\bigskip Recently two of the authors have proposed~\\cite{BF,F,BF2} a scenario for a cyclic universe based on a dark energy component with constant equation of state satisfying $w = p/\\rho = -1-\\phi$ where $w < -1$ and hence $\\phi > 0$. The model involves two key ideas: (i) that the universe deflate just prior to the turnaround from expansion to contraction by disintegrating into a very large number $N_{\\rm cp}$ of causal patches. In the notation of \\cite{BF}, note that $N_{\\rm cp}=1/f^3$; (ii) that the contracting universe be empty, meaning that one causal patch at turnaround must contain no matter or black holes, only dark energy. This is called the CBE condition ({\\it Comes Back Empty}). Implementation of CBE requires, as we shall explain, a lower bound on $N_{cp}$ which depends on the length scale $L$ characterising the smallest bound system. We shall consider both $L = 10^{-15}$m for a nucleon then $L \\ge 10^{-35}$m for a PPP ({\\it Presently Point Particle}) meaning a particle which at present is considered to be pointlike, like a quark or a lepton, but which actually has a characteristic size at least a few orders of magnitude smaller than a nucleon but greater than or equal to the Planck scale. \\bigskip In the foreseeable future, it is expected that the equation of state of the dark energy $w$, and hence $\\phi$, will be measured with higher accuracy by, for example, the Planck Surveyor satellite~\\cite{Planck}. What we shall show is that this measurement can, within this model, constrain for a given $w$ the number $N_{\\rm cp}$ of causal patches at turnaround by imposing a lower bound thereon. \\bigskip The plan of the paper is that in Section 2 we discuss the times at which unbinding, causal disconnection and turnaround occur. In Section 3, the constraints on $N_{\\rm cp}$ from measurement of $\\phi$ are derived. Finally, Section 4 is a discussion. In the Appendices is technical material to supplement the main text. \\newpage ", "conclusions": "\\bigskip What we have deduced is that the parameters $\\rho_c, w$ and $N_{\\rm cp}$ in cyclic cosmology are already constrained by existing data. For example one requires $w \\gtrsim -2$ for the CBE aspect to work. \\bigskip This constraint is already known to be respected in Nature but as better and more accurate cosmological data become available it will shed further light on the viability of the theory. \\bigskip In particular, the accurate measurement of the equation of state $w = -1 -\\phi$ is of special interest. Fortunately the Planck Surveyor~\\cite{Planck} is anticipated to acquire improved accuracy on $w$ in the near future. As we have discussed, this will provide a lower bound on the number $N_{\\rm cp}$ of causal patches necessary to dissociate the smallest bound systems at turnaround and hence to solve the entropy problem and, via CBE, enable the possibility of infinite cyclicity. \\bigskip It is amusing that the physical conditions at the approach of deflation are so extraordinary that it is natural to ask whether the systems presently regarded as point particles may be composite because the phantom dark energy density grows to unimaginably large values and can disintegrate bound systems down to arbitrarily small scales. We have conservatively limited our attention to systems bigger than the Planck length. However, although this requirement seems dictated by considerations of quantum gravity, it is possible that the dark energy will dissociate even smaller systems if they exist. \\bigskip The advantage of cyclic cosmology is that it removes the initial singularity associated with the Big Bang, about 13.7 billion years ago, and allows that time never began. The previous attempts to create a consistent infinite cyclicity were stymied between about 1934~\\cite{Tolman} and 2002~\\cite{ST} primarily because of the entropy problem and the second law of thermodynamics. The discovery of the accelerated expansion rate of the universe and the concomitant necessity of dark energy has permitted more optimism that the cyclic cosmology is, after all, on the right track. \\newpage \\begin{center}" }, "0801/0801.3036_arXiv.txt": { "abstract": " ", "introduction": "\\label{intro} As we move towards the era of ELTs, it is timely to think about the future role of the 8-m class telescopes. Under the OPTICON programme novel technologies have been developed that are intended for use in multi-object and integral-field spectrographs. To date, these have been targeted at instrument concepts for the European ELT, but there are also significant possibilities for their inclusion in new VLT instruments, ensuring the continued success and productivity of these unique telescopes. ", "conclusions": "In summary, there are considerable technical challenges in developing a wide-field spectrograph for the VLT, particularly when one considers the available resources within ESO and of its partners; a wide-field prime focus instrument would also lead to operational problems for VLTI. Future smart focal plane and photonics technologies may enable a lighter, more compact, prime focus (or foward-Cassegrain) instrument, but current technology readiness levels suggest that a compromise would be to exploit the maximum field available at the Nasmyth foci, e.g. Super-Giraffe \\cite{giraffe}. Meanwhile, technology development for ELT instrument concepts such as EAGLE, and through the next Framework Programme, should be supported to help break the existing paradigm. \\vspace{-0.1in}" }, "0801/0801.1205_arXiv.txt": { "abstract": "{} {We revisit the vicinity of the microquasar Cygnus X-3 at radio wavelengths. We aim to improve our previous search for possible associated extended radio features/hot spots in the position angle of the Cygnus X-3 relativistic jets focusing on shorter angular scales than previously explored. } {Our work is mostly based on analyzing modern survey and archive radio data, mainly including observations carried out with the Very Large Array and the Ryle Telescopes. We also used deep near-infrared images that we obtained in 2005. } {We present new radio maps of the Cygnus X-3 field computed after combining multi-configuration Very Large Array archive data at 6 cm and different observing runs at 2 cm with the Ryle Telescope. These are probably among the deepest radio images of Cygnus X-3 reported to date at cm wavelengths. Both interferometers reveal an extended radio feature within a few arc-minutes of the microquasar position, thus making our detection more credible. Moreover, this extended emission is possibly non-thermal, although this point still needs confirmation. Its physical connection with the microquasar is tentatively considered under different physical scenarios. We also report on the serendipitous discovery of a likely Fanaroff-Riley type II radio galaxy only $3^{\\prime}$ away from Cygnus X-3. } {} ", "introduction": "Cygnus X-3 is one of the X-ray binaries considered as the prototype of the family of galactic sources with relativistic jets, i.e., microquasars. It was originally discovered in X-rays by \\cite{gia67}. This system made headlines more than three decades ago when its first strong radio outbursts were detected, as described in \\cite{g72} and subsequent papers. Such flaring events have been repeating since then, typically one or two times per year, with flux density increments of almost two orders of magnitude above the normal quiescent level of $\\sim0.1$ Jy at cm wavelengths (\\cite{w1994}). At present, Cygnus X-3 is considered to be a high-mass X-ray binary (HMXB) with a WN Wolf-Rayet (WR) companion star of WN8 type (see e.g. \\cite{ke1996} and \\cite{ko2002}). The nature of the compact object is not well constrained, and neither a neutron star nor a black hole accreting from the WR star can be ruled out (see e.g. \\cite{er1998} and \\cite{mi1998}). The observed X-ray (\\cite{p1972}) and infrared (\\cite{bk1973}) modulation of 4.8 h is believed to be connected with the system orbital period, which is rather short for an HMXB. Orbital-phase-resolved spectra of Cygnus X-3 in the near infrared are consistent with it (\\cite{han2000}). The distance has been estimated to be approximately 9 kpc (\\cite{pred00}), which agrees with a strong interstellar absorption ($A_V \\sim 20$ mag, \\cite{ke1996}) that renders the optical counterpart undetectable in the visual domain. \\begin{table*} \\begin{center} \\caption[]{\\label{vlaobs} VLA archive observations used in this paper \\label{obslog}} \\begin{tabular}{cccccccc} \\hline \\hline Date & wavelength & VLA & Number of & Bandwidth & Number of & Project & Time on \\\\ & (cm) & configuration. & IFs & (MHz) & visibilities & id. & source (hrs) \\\\ \\hline 1992 Jun 8 & 6 & DnC & 2 & 50 & 432229 & UT002 & 5 \\\\ 1997 May 4 & 6 & B & 2 & 50 & 718939 & AM551 & 5.7 \\\\ 2000 Aug 29 & 6 & D & 2 & 50 & 29167 & AH669 & 0.33 \\\\ \\hline \\hline \\end{tabular} \\end{center} \\end{table*} Galactic microquasars such as Cygnus X-3 are known to release a significant amount of energy in the form of collimated relativistic jets into their surrounding inter-stellar medium (ISM). For Cygnus X-3, the averaged energy injection rate of its relativistic jets is estimated to be at least $10^{37}$ erg s$^{-1}$ (\\cite{m2005}), i.e., enough to supply $\\sim 10^{50}$ erg (10\\% of a supernova explosion) during the expected lifetime of the WR star. The Cygnus X-3 jets have been repeatedly resolved as sub-parsec transient radio features, propagating at a significant fraction of the speed of light in the north-south direction, thanks to interferometric radio techniques from arc-second (\\cite{m2001}) to milli arc-second angular scales (e.g. \\cite{g1983,miller04,tu2007a}). However, up to now there is no robust evidence of interaction between the relativistic ejecta of the system and the surrounding ISM on larger, few pc scales. This is in contrast with other microquasars, such as Cygnus X-1 or Circinus X-1, where a clear signature of interaction between their relativistic jets and the ISM does exist with a ring-like or lobe morphology several pc wide (\\cite{ga2005,tu2007b}). Additional examples of relativistic jet/ISM interaction come from the also pc-scale lobes of the Galactic Centre Annihilator 1E1740.7$-$2942 (\\cite{mira92}) or GRS 1758$-$258 (\\cite{m2002}). The detection of such ring/lobe features is a useful tool to better constrain the system's energetics. The search for large scales features associated with the Cygnus X-3 radio jets has been a concern of the authors during recent years. Two H\\ion{II}\\ regions located $\\sim40^{\\prime}$ from Cygnus X-3 along the jet position angle were first tentatively proposed as possible large scale lobes (\\cite{m2000}). However, no further evidence for a physical connection could be found beyond the mere geometric alignment. A closer search revealed later the existence of two possible hot spot candidates (HSCs) associated with Cygnus X-3, thus suggesting an analogy with Fanaroff-Riley type II (FR II) radio galaxies (\\cite{m2005}). The apparent hot spots were two faint radio sources with non-thermal spectra at angular distances of 7\\prp 07 and 4\\prp 36 from Cygnus X-3. The line joining them was also within one degre of the almost North-South position angle of the inner arc-second radio jets (\\cite{m2001}). Unfortunately, follow up radio and near infrared observations of both the HSCs and the Cygnus X-3 nearby environment did not confirm the proposed hot spot nature and indicated that they were most likely background or foreground objects (\\cite{m2006}). This fact left open again for Cygnus X-3 the issue of searching for signatures of energy deposition from its relativistic jets into the ISM. In this context, the main purpose of this paper is to present new very deep radio images of Cygnus X-3 and its environment obtained after combining multi-epoch archive and survey data, together with our own observations. The resulting maps enable us to put very strong limits for any extended or compact radio features that could be associated to Cygnus X-3 on a scale of a few pc. We also report on several extended radio features in the field with apparent non-thermal spectra that were previously unknown. Their possible connection with the microquasar is discussed from a skeptical point of view. ", "conclusions": "\\begin{enumerate} \\item We revisited the issue of large scale radio features associated to the microquasar Cygnus X-3. By combining different archive VLA observations, we have been able to create a very deep (rms noise 9.5 $\\mu$Jy beam$^{-1}$) radio map with sensitivity to arc-minute angular scales. Cygnus X-3 appears superposed onto a diffuse radio emission with apparent non-thermal index with an angular size of a few arc-minutes extending South and South-West from it. The reality of such feature is independently confirmed by Ryle Telescope observations and when inspecting in detail maps from the CGPS and other survey data. \\item With all cautions in mind, we tentatively suggest the possibility that such an extended emission could be physically associated to Cygnus X-3. It could be either a distorted lobe, plume or partial shell-like structure resulting from the accumulated flaring history of the microquasar interacting with the ISM. If connected with Cygnus X-3, the feature linear size would be $\\sim5$ pc. This is of the same order as extended lobes or bubbles seen around other microquasars. Based on simple equipartition assumptions, the observed size is roughly in agreement with expectations from simple theory of radio lobe dynamics. \\item A triple radio source with FR II morphology has been also discovered at few arc-minute from Cygnus X-3 in our deep VLA map. In addition to provide a curious `family picture' of two accreting sources in the same shot, the angular proximity of the FR II compact core could enable future proper motion studies of this microquasar using future generations of highly sensitive interferometers. \\end{enumerate}" }, "0801/0801.3941_arXiv.txt": { "abstract": "The dynamical mass of a star cluster can be derived from the virial theorem, using the measured half-mass radius and line-of-sight velocity dispersion of the cluster. However, this dynamical mass may be a significant overestimation of the cluster mass if the contribution of the binary orbital motion is not taken into account. In these proceedings we describe the mass overestimation as a function of cluster properties and binary population properties, and briefly touch the issue of selection effects. We find that for clusters with a {\\em measured} velocity dispersion of $\\sigmalos\\ga 10$~\\kms{} the presence of binaries does not affect the dynamical mass significantly. For clusters with $\\sigmalos\\la 1$~\\kms{} (i.e., low-density clusters), the contribution of binaries to $\\sigmalos$ is significant, and may result in a major dynamical mass overestimation. The presence of binaries may introduce a downward shift of $\\Delta \\log (L_V/\\mdyn) = 0.05-0.4$ in the $\\log (L_V/\\mdyn)$ vs. age diagram. ", "introduction": "% An estimate for the mass of star clusters can be obtained from the virial theorem, using the projected half-mass radius $\\halfmass$ and line-of-sight velocity dispersion $\\sigmalos$. This dynamical mass estimate, $\\mdyn$, is obtained using the equation derived by \\cite{spitzer1987}: \\begin{equation} \\label{equation:spitzer} \\mdyn = \\eta \\, \\frac{\\halfmass \\sigmalos^2}{G} \\,, \\end{equation} where $\\eta \\approx 9.75$ is a dimensionless proportionality constant. The derivation of $\\mdyn$ using the expression above is valid under the following assumptions: (1) the cluster dynamics are described by the Plummer model, (2) all stars are single and of equal mass, (3) the cluster is in virial equilibrium, and (4) no selection effects are present. Dynamical mass estimates for numerous clusters have been obtained this way \\citep[e.g.,][]{mandushev1991,smith2001,maraston2004,bastian2006,larsen2007,moll2007}. When binary stars are present, however, Eq.~(\\ref{equation:spitzer}) results in an overestimation of the cluster mass. Observations have shown that the majority of the field stars are part of a binary or multiple system \\citep{duquennoy1991,fischermarcy1992}. It is believed that most (if not all) binary stars are formed in binary systems, which is supported by both observations \\citep{mathieu1994,mason1998,kobulnicky2007,kouwenhoven_adonis,kouwenhoven_recovery} and theory \\citep{goodwinkroupa2005}. Stars in binary systems exhibit not only motion in the gravitational potential of the cluster ({\\em particle motion} $\\sigmapart$), but also {\\em orbital motion} $\\sigmaorb$ in the binary system. Eq.~(\\ref{equation:spitzer}) is applicable for $\\sigmapart$ (the centre-of-mass motion of the binaries), but results in an overestimation if $\\sigmalos$, the superposition of $\\sigmapart$ and $\\sigmaorb$, is measured. In this paper we therefore address the question: ``How do binaries affect the dynamical mass of a star cluster?'' ", "conclusions": "% The total mass of a star cluster is often inferred from its line-of-sight velocity dispersion $\\sigmalos$ and half-mass radius $\\halfmass$, assuming virial equilibrium. The latter approach includes the assumption that no binaries are present. However, most stars are known to be in binary systems, and their orbital motions provide an additional contribution to the measured $\\sigmalos$. The latter value is now no longer representative for the (centre-of-mass) motion of the stars and binaries in the cluster potential. This effect may result in a significant dynamical mass overestimation. Depending on the magnitude of the dynamical mass overestimation, we distinguish between three types of clusters: particle-dominated ($\\sigmaorb \\ll \\sigmapart$), intermediate-case ($\\sigmaorb \\approx \\sigmapart$), and binary-dominated clusters ($\\sigmaorb \\gg \\sigmapart$). The orbital velocity of binaries is independent of the cluster properties (size, mass, etc.). Whether or not binary motion affects $\\sigmalos$ is thus depends on the cluster properties. For clusters with a high stellar density (i.e, large $\\mcl$ or small $\\halfmass$), $\\mdyn$ is generally unaffected. In the latter case $\\sigmapart > 10$~\\kms{}. The dynamical mass overestimation increases strongly with (i) a higher binary fraction and (ii) a smaller average orbital size. The dependence on the mass ratio distribution and eccentricity distribution is small: $\\Delta(\\mdyn/\\mcl) \\la 5\\%$. We additionally show that observing a certain subset of the cluster introduces a selection effects, which may result in a further mass overestimation by up to 40\\%. Analysis of the brightest (generally massive) stars in the cluster may further overestimate the dynamical mass. The dynamical mass overestimation can be negligible for the most massive clusters, while it may overestimate the true mass by up to an order of magnitude for sparse OB associations. The full results of this study have been presented in \\cite{kouwenhoven_sigma}. A follow-up paper, which treats the selection effects properly, is in preparation." }, "0801/0801.3988_arXiv.txt": { "abstract": "V5116 Sgr (Nova Sgr 2005 No. 2), discovered on 2005 July 4, was observed with XMM-Newton in March 2007, 20 months after the optical outburst. The X-ray spectrum shows that the nova had evolved to a pure supersoft X-ray source, with no significant emission at energies above 1~keV. The X-ray light-curve shows abrupt decreases and increases of the flux by a factor $\\sim8$. It is consistent with a periodicity of 2.97~h, the orbital period suggested by \\cite{dob07}, although the observation lasted just a little more than a whole period. We estimate the distance to V5116~Sgr to be $11\\pm3\\,kpc$. A simple blackbody model does not fit correctly the EPIC spectra, with $\\chi^2_{\\nu}>4$. In contrast, ONe rich white dwarf atmosphere models provide a good fit, with $N_{\\rm H}= 1.3(\\pm0.1)\\times 10^{21}\\,cm^{-2}$, $T=6.1(\\pm0.1)\\times10^5\\,K$, and $L=3.9(\\pm0.8)\\times10^{37}(D/10~\\rm{kpc})^2{\\rm erg}\\,{\\rm s}^{-1}$ (during the high-flux periods). This is consistent with residual hydrogen burning in the white dwarf envelope. The white dwarf atmosphere temperature is the same both in the low and the high flux periods, ruling out an intrinsic variation of the X-ray source as the origin of the flux changes. We speculate that the X-ray light-curve may result from a partial coverage by an asymmetric accretion disk in a high inclination system. ", "introduction": "Classical novae occur in close binary systems of the cataclysmic variable type, when a thermonuclear runaway results in explosive hydrogen burning on the accreting white dwarf (WD). It is theoretically predicted that novae return to hydrostatic equilibrium after the ejection of a fraction of the accreted envelope. Soft X-ray emission arises in some post-outburst novae as a consequence of residual hydrogen burning on top of the WD. As the envelope mass is depleted, the photospheric radius decreases at constant bolometric luminosity (close to the Eddington value) with an increasing effective temperature. This leads to a hardening of the spectrum from optical through UV, extreme UV and finally soft X-rays, with the post-outburst nova emitting as a Supersoft Source (SSS) with a hot WD atmosphere spectrum. The duration of this SSS is related to the nuclear burning timescale of the remaining H-rich envelope and depends, among other factors, on the WD mass \\citep{TT98,SH05a,SH05b}. All post-outburst novae are expected to undergo SSS emitting phase, provided a H-rich envelope is left. Nevertheless, the number and duration of SSS states observed in Galactic novae is small. In a systematic search for the X-ray emission from novae in the ROSAT archive, \\cite{Ori01a} found only three SSS novae from a total of 39 novae observed less than ten years after the outburst (GQ Mus, V1974~Cyg, N~LMC~1995). After the ROSAT era, observations by Beppo-SAX and Chandra revealed SSS X-ray emission for some more post-outburst novae: V382~Vel \\citep{Ori02,Bur02}, V1494~Aql, with a puzzling light curve showing a short burst and oscillations \\citep{Dra03}, and V4743~Sgr, also with a very variable light-curve \\citep{Nes03,Pet05}. The SSS phase of the recurrent nova RS~Oph in the 2006 outburst was monitored with Swift, XMM-Newton and Chandra, and the end of the SSS state occurred less than 100 days after outburst \\citep{Bode06,Hachi07}. XMM-Newton has contributed with some more observations of novae. Nova~LMC 1995 showed the SSS emission 5 years after outburst \\citep{Ori03}. Other post-outburst novae were detected with a harder spectrum associated to the expanding ejecta, as Nova LMC 2000 \\citep{Gre03} and V4633 Sgr \\citep{HS07}, or reestablished accretion flow \\cite[V2487~Oph,][]{HS02,fer07}. More recently, the SSS XMMSL1~J070542.7-381442 was identified as a nova \\citep{rea07a,rea07b,tor07}. X-ray observations of the central area of M31, with its moderate absorption, offer a good chance to monitor the SSS phase of novae: \\cite{pie05} reported 21 X-ray counterparts for novae in M31 -- mostly identified as SSS by their hardness ratios -- and two in M33. Following that work, XMM-Newton and Chandra monitoring of M31 between July 2004 and February 2005 provided the detection of eleven out of 34 novae within a year after optical outburst \\citep{pie07}. This suggests that the low fraction of SSS found by \\cite{Ori01a} was due to selection effects and/or too poor sampling. X-ray observations of post-outburst novae provide crucial information: soft X-rays yield a unique insight into the remaining burning envelope on top of the WD, while hard X-rays reveal shocks in the nova ejecta. The properties of ``quiescent novae'', once they have turned-off and once accretion is reestablished (which can occur before or after the turn-off), are revealed both in hard and soft X-rays; they emit then as ``standard'' cataclysmic variables. In view of the scarcity of objects observed and the diversity of behaviors detected, only the monitoring of as many novae as possible, with large sensitivity and spectral resolution (as those offered by XMM-Newton and Chandra) can help to understand post-outburst novae. V5116 Sgr (Nova Sgr 2005 No. 2) was one of the targets included in our X-ray monitoring programme of post-outburst Galactic novae with XMM-Newton. It was discovered by \\cite{lil05} on 2005 July 4.049~UT, with magnitude $\\sim$8.0, rising to mag 7.2 on July 5.085. The expansion velocity derived from a sharp P-Cyg profile detected in a spectrum taken on July 5.099 was $\\sim$1300~km/s. IR spectroscopy on July 15 showed emission lines with FWHM~$\\sim$2200~km/s \\citep{rus05}. Photometric observations obtained during 13 nights in the period August-October 2006 show the optical light curve modulated with a period of $2.9712\\pm0.0024$~h \\citep{dob07}, which the authors interpret as the orbital period. They propose that the light-curve indicates a high inclination system with an irradiation effect on the secondary star. A first X-ray observation with Swift/XRT (0.3--10~keV) in August 2005 yielded a marginal detection with 1.2($\\pm$1.0)$\\times10^{-3}$~cts/s \\citep{nes07a}. Two years later, on 2007 August 7, Swift/XRT showed the nova as a SSS, with 0.56$\\pm$0.1~cts/s \\citep{nes07b}. A first fit with a blackbody indicated $T\\sim4.5\\times10^5K$. More recently, a 35~ks Chandra spectrum obtained on 2007 August 28 was fit with a WD atmospheric model with N$_{\\rm H}=4.3\\times10^{21}$~cm$^2$ and $T=4.65\\times10^5K$ \\citep{NelOri07}. Here we report on the X-ray light-curve and broad-band spectra of V5116 Sgr as observed by XMM-Newton in March 2007, 609 days after outburst. ", "conclusions": "V5116 Sgr was detected as a bright SSS by XMM-Newton 20 months after the nova outburst. The temperature of the WD atmosphere indicates that residual hydrogen burning is occurring in the WD envelope. To further confirm the origin of the emission and determine the size of the emitting surface, we need an estimation of the distance. Fortunately enough, the time and magnitude at maximum are quite well determined for V5116~Sgr, with a pre-maximum observation only 24~hours before the maximum. \\cite{lil05} reported $m_V=7.15$ at maximum, and $t_{2}$ (time required to decline two magnitudes from maximum) was 6.5$\\pm$1.0~days \\citep{dob07}. Using the empirical relation for novae between $M_V$ and $t_2$ \\citep{del95} the observed $t_2$ implies an absolute magnitude at maximum $M_V=-8.8(\\pm0.4)$ (we add a 5\\% error to account for the scatter in the $M_V-t_2$ relation). The photometry at maximum indicates $B-V=+0.48$ \\citep{gil05}, and for novae at maximum, intrinsic $B-V=0.23\\pm0.06$ \\citep{ber87}. This implies $A_{V}=3.1\\,E_{B-V}=0.8\\pm0.2$. With all this, we estimate the distance to V5116~Sgr to be $d=11\\pm3\\,kpc$. For simplicity, we scale the distance dependent parameters to 10~kpc. It is worth noticing that the $N_{\\rm H}$ obtained from our fits indicates $A_{V}=0.7$ \\cite[$N_{\\rm H}=5.9\\times10^{21}E_{B-V}\\,\\mbox{cm}^{-2}$,][]{zom07}, consistent with the value obtained from the observed colors. The luminosity during the high flux periods, together with the atmosphere temperature, indicate a radius of the emitting object $R=6.2(\\pm0.9)\\times10^8(D/10kpc)\\,cm$. This corresponds to the whole WD surface emitting during the high flux periods and supports residual H-burning as the origin of the SSS emission, rather than any feature related to the accretion stream (boundary layer, hot spot, or hot accretion poles). The most remarkable feature in V5116~Sgr is the SSS light-curve. Although the observation lasted just a bit longer than an orbital period, we obtain a period compatible with the orbital period of 2.97~hours found by \\cite{dob07}. The fact that the temperature is the same for both the high and the low flux phases indicates that the change in flux cannot be caused by an intrinsic change in the H-burning envelope. The luminosity during the high flux periods also indicates that the whole WD is visible in this state. The decrease by a factor $\\sim8$ in the flux could be caused by a partial eclipse of the WD. The long duration of the low flux phase and short duration of the high flux phase cannot be reconciled with a partial eclipse by the secondary star. Alternatively, some asymmetric accretion disk could do the job, being responsible for a partial eclipse during most of the orbital period, with a 3~ks window during which the whole WD would be visible. Since no change is observed in the hydrogen absorption column and it is compatible with the average interstellar value both in the low and the high flux spectra, the portion of accretion disk producing the partial eclipse should be optically thick to the soft X-rays. In the particular case of V5116 Sgr, we have checked that an absorbing column of $N_{H}>2\\times10^{22}\\,cm^{-2}$ would make it completely opaque to the observed SSS emission. Longer X-ray observations covering at least two periods together with a good timing with the OM camera would be required to disentangle the origin of the X-ray light curve behaviour." }, "0801/0801.1496_arXiv.txt": { "abstract": "We report high-cadence time-series photometry of the recently-discovered transiting exoplanet system HD~17156, spanning the time of transit on UT 2007 October 1, from three separate observatories. We present a joint analysis of our photometry, previously published radial velocity measurements, and times of transit center for 3 additional events. Adopting the spectroscopically-determined values and uncertainties for the stellar mass and radius, we estimate a planet radius of $R_{p} = 1.01 \\pm 0.09 \\, R_{\\rm Jup}$ and an inclination of $i = 86.5^{+1.1}_{-0.7}$~degrees. We find a time of transit center of $T_{c} = 2454374.8338 \\pm 0.0020$~HJD and an orbital period of $P = 21.21691 \\pm 0.00071$~days, and note that the 4 transits reported to date show no sign of timing variations that would indicate the presence of a third body in the system. Our results do not preclude the existence of a secondary eclipse, but imply there is only a 9.2\\% chance for this to be present, and an even lower probability (6.9\\%) that the secondary eclipse would be a non-grazing event. Due to its eccentric orbit and long period, HD~17156b is a fascinating object for the study of the dynamics of exoplanet atmospheres. To aid such future studies, we present theoretical light curves for the variable infrared emission from the visible hemisphere of the planet throughout its orbit. ", "introduction": "\\label{intro_sect} Twenty-eight transiting planets are now reported in the literature\\footnote{See {\\tt http://www.inscience.ch/transits/}}, and the doubling time scale for new detections is now roughly one year. As reviewed by \\citet{char2007}, it is these objects that have allowed the study of the physical structure and atmospheric chemistry and dynamics of gas giant exoplanets, and opened the field of comparative exoplanetology. However, the majority have orbital periods of a few days, due both to the lower geometric probability of transits occurring as the orbital period increases, and due to the limitations of ground-based photometric transit surveys that strongly favor detection of short-period systems. The latter problem can be circumvented by searching for transits of known radial velocity planet-bearing stars, putting to use the radial velocity information to constrain the possible time of transit. This is the primary aim of the Transitsearch.org network \\citep{sea2003}\\footnote{See {\\tt http://www.transitsearch.org}}, which, starting in 2002, has been conducting photometric follow-up observations of known radial velocity detected planets. This network is a prime example of successful collaboration between amateur and professional astronomers. Although longer period planets possess a lower geometric probability to transit, there exists a loop-hole for some planets on highly eccentric orbits. The transit probability for a planet on an eccentric orbit, with periastron near inferior conjunction, is amplified by a factor \\begin{equation} A = \\left[ {{1+e\\cos(\\frac{\\pi}{2} - \\omega)} \\over{1-e^2}}\\right] \\end{equation} where $e$ is the eccentricity, and $\\omega$ is the argument of pericenter \\citep{sea2003}. An amplification factor near 2.9 for the planet orbiting HD~17156 \\citep{fischer2007}, along with an extremely short time window for possible transits, brought the system to the attention of the Transitsearch.org network. This was rewarded by the detection of transits as recently announced by \\citet{bar2007}. This unique system contains a $3.1$-$M_{\\rm Jup}$ transiting planet in a $21.2$ day highly-eccentric orbit ($e = 0.67$) about a bright ($V = 8.2$) G0V star. \\citet{gil2007} have recently presented refined estimates of the system parameters based on new photometric data. Even among planets with periods beyond 5 days, where the timescale for tidal circularization quickly grows to exceed the age of most systems, HD~17156b's eccentricity is unusually high, thus making it an interesting test case for models of planetary migration. Preliminary measurements of the Rossiter-McLaughlin effect \\citep{rossiter24,mclaughlin24,gw2007,winnasp2007} by \\cite{nar2007} show evidence for a misalignment between the stellar spin axis and the planetary orbit axis, which may indicate a migration mechanism involving planet-planet scattering \\citep[see e.g.][]{chat2007}, or Kozai migration under perturbation by a yet undetected stellar companion or second planet \\citep{fab2007,wu2007}. The large eccentricity and long period also make HD~17156 a particularly attractive target for several additional follow-up studies. First, it presents a unique opportunity for the study of the structure and dynamics of a gas-giant atmosphere under strongly varying illumination, through infrared monitoring of the planetary emission (e.g. \\citealt{har2006}; \\citealt{cow2007}; \\citealt{knut2007a,knut2007b}). Second, by monitoring successive times of transit, the presence of additional bodies in the system can be detected or constrained from the resulting perturbations on the orbit of HD~17156b (\\citealt{hm2005}; \\citealt{agol2005}; \\citealt{sa2005}). The purpose of our paper is three-fold: First, we seek to refine the estimates of the system parameters that were only poorly constrained by the light curves of \\citet{bar2007}. Second, we document the time of center of transit and search for transit timing variations. Third, we evaluate the likelihood that a secondary eclipse will be observable for HD~17156, and present theoretical predictions of the infrared phase variations as might be detected with the {\\it Spitzer Space Telescope}. We elected to perform an independent analysis from that of \\citet{gil2007}, and therefore do not use their orbital parameters as constraints, since at the time of writing these are still subject to change as their paper has not yet been accepted for publication. We do however make use of the time of mid-transit presented in their Table 1. We begin by presenting a description of our observations of the transit event on UT 2007 October 1 in \\S \\ref{obs_sect}. We then present our combined analysis of the extant radial velocity measurements (\\S \\ref{rv_sect}) and photometric data (\\S \\ref{phot_sect}), and our resulting estimates of the system parameters and likelihood of a secondary eclipse. In \\S \\ref{disc_sect}, we compare our constraints on the planet mass and radius with planetary structural models, summarize the current constraints on the presence of transit timing variations in the system, and present model predictions for the planetary infrared light curve. We conclude in \\S \\ref{summ_sect} with a discussion of compelling avenues for future research. ", "conclusions": "\\label{summ_sect} HD~17156 thus represents a highly interesting system for follow-up studies in the mid infra-red (e.g. using the {\\it Spitzer Space Telescope}), even if it does not undergo secondary eclipses. Ground-based observations should also be pursued to continue the search for transit timing variations, and hence to constrain the presence of additional bodies in the system, which, given the high orbital eccentricity of the planet HD~17156b, may show interesting dynamical properties and place constraints on planetary formation and evolution scenarios. This discovery represents an important success for the Transitsearch.org project, and highlights the potential rewards of collaboration between distributed networks of amateur astronomers and the professional community." }, "0801/0801.4085_arXiv.txt": { "abstract": "We use several main-sequence models to derive distances (and extinctions), with statistically meaningful uncertainties for 11 star-forming-regions and young clusters. The model dependency is shown to be small, allowing us to adopt the distances derived using one model. Using these distances we have revised the age order for some of the clusters of \\cite{2007MNRAS.375.1220M}. The new nominal ages are: $\\approx2$ Myrs for NGC6530 and the ONC, $\\approx3$ Myrs for $\\lambda$ Orionis, NGC2264 and $\\sigma$ Orionis, $\\approx4-5$ Myrs for NGC2362, $\\approx13$ Myrs for h and $\\chi$ Per, $\\approx20$ Myrs for NGC1960 and $\\approx40$ Myrs for NGC2547. In cases of significantly variable extinction we have derived individual extinctions using a revised Q-method \\citep{1953ApJ...117..313J}. These new data show that the largest remaining uncertainty in deriving an age ordering (and necessarily ages) is metallicity. We also discuss the use of a feature we term the R-C gap overlap to provide a diagnostic of \\textbf{isochronal} age spreads or varying accretion histories within a given star-formation-region. Finally, recent derivations of the distance to the ONC lie in two groups. Our new more precise distance of $391^{+12}_{-9}$ pc allows us to decisively reject the further distance, we adopt 400 pc as a convenient value. ", "introduction": "\\label{intro} Colour-magnitude diagrams (CMDs) of star-formation regions (SFRs) provide, in combination with model isochrones, an excellent tool with which to determine distances, ages and individual stellar masses. These parameters are critical for determining initial mass functions (IMFs) for stellar populations and discovering the possible impacts of local environment (such as the effect of ionising winds from massive stars) on disc lifetimes and on star and planet formation and evolution. Many calculations of IMFs, disc fractions etc are available but they are derived in heterogeneous ways, thus hints of the effects of environment are only recently beginning to emerge \\citep[e.g.][]{2007MNRAS.375.1220M,2004AJ....128..765S}. Very precise photometry ($\\approx1\\%$) is routinely available, along with sophisticated stellar models. However, current parameter derivations from CMDs still have relatively large uncertainties and are model dependent \\citep[see the discussions in][]{2004A&A...415..571B, 2004ApJ...600..946P,2002MNRAS.335..291N,2007MNRAS.375.1220M}. Thus current age (and distance) uncertainties all but `wash-out' any environmental effects. Clearly more robust constraints would be available for current stellar theories if more precise parameters could be extracted from the CMDs of SFRs. In \\cite{2007MNRAS.375.1220M} we created an age ladder for a range of pre-MS populations. The first stage was to create empirical isochrones by fitting splines to the pre-MS locus. Overlaying them in absolute magnitude and intrinsic colour results in an age ladder, with the youngest SFRs at the brightest absolute magnitudes. SFRs with almost indistinguishable positions in the CMD were grouped, and nominal ages assigned to each group. Thus the age sequence (though not the nominal ages) are free from the problems associated with pre-main-sequence (pre-MS) models. In \\cite{2007MNRAS.375.1220M} we had to adopt literature distances for the studied SFRs. These distances were derived using a range of different methods and their uncertainties proved to be the largest remaining contributor to the uncertainties in our age ladder placements. Previous distances have chiefly been derived using main-sequence (MS) isochrone fitting, pre-MS isochrone fitting or from \\textit{HIPPARCOS} parallax measurements. MS isochrone fitting provides distances based on the positions of MS stars in a CMD, which are independent of uncertainties in age. Pre-MS isochrone fitting also uses the positions of stars in a CMD, but in this method the derived distances are degenerate with age \\citep[see e.g.][]{2006MNRAS.373.1251N}. Finally, distances derived from \\textit{HIPPARCOS} parallax measurements are only available for a few SFRs included in this paper, $\\sigma$ Ori, NGC2547 and $\\lambda$ Ori, with all except $\\lambda$ Ori having large uncertainties. Of those methods used to derive distances, the most suitable for the derivation of an age ladder is clearly MS isochrone fitting. Fully convective pre-MS stars (in young SFRs) are separated in a CMD from those stars on the MS which have radiative cores. The transistion region or gap in the CMD (measurable in colour), we term the radiative-convective gap \\citep[R-C gap, see][for introduction of term and discussion]{2007MNRAS.375.1220M}. Once stars have crossed the R-C gap their position in a CMD is almost independent of age until they reach the turn off. However MS fitting has not yielded the precision one would expect in distance estimates. This is due to two significant problems. Firstly the position of MS isochrones, although temporally static, is model dependent, with different studies adopting different models. Secondly distances are most often derived using `by eye' fitting of models to the data, yielding ill-defined uncertainties. A full discussion of previous fitting methods can be found in \\cite{2006MNRAS.373.1251N}. In this paper we solve both of these problems. Firstly we show that the model dependency is small for the model isochrones studied. We then adopt the distances from a single MS model. This allows us to derive a set of precise distances, accurate relative to each other, which we term \"relative distances\". Second we use the $\\tau^2$ fitting technique \\citep{2006MNRAS.373.1251N}, a new rigorous and self-consistent method of fitting stars to isochrones, which yields statistically meaningful uncertainties. This presents us with the opportunity to achieve more precise distances from the fitting of high-mass (HM) or MS stars. The rest of this paper is laid out as follows. In Section \\ref{data} we detail the literature sources, the nature of the data used and any sequence selection carried out on the stars. Section \\ref{calibration} details the different model isochrones and photometric calibrations used. Section \\ref{fit_method} describes the fitting process. This is done primarily by way of an example in Section \\ref{chiper_eg}. Section \\ref{ind_ext} describes the derivation of individual extinctions, in particular Section \\ref{Q} describes an revised Q-method for calculating approximate individual extinctions. The results for all the isochrone calibrations and methods are presented in Section \\ref{results_app}. Section \\ref{results} outlines our results for one adopted model with our best-fitting distances given in Table \\ref{dist_comp}. In section \\ref{implications} we discuss the implications of the individual distances to several key SFRs (Section \\ref{imp_ind}) and those of the entire dataset. The implications of the dataset on metallicity (Section \\ref{imp_met}), age spreads and the R-C gap overlap (Section \\ref{imp_cuts}) and secular evolution within the SFRs with particular reference to disc fractions (Section \\ref{imp_sec}) are discussed. The reader interested in distance and reddening values should skip to Section \\ref{results_app} for the values derived from all the models used. ", "conclusions": "\\label{imp_sum} \\begin{enumerate} \\item We have derived a self-consistent set of distances of generally higher precision than previously available for a set of SFRs in the age range 1-40 Myrs. We have also derived distances using several other models and calibrations (see Section \\ref{results_app}). \\item In addition to these new distances and reddenings (or extinctions) we have reconstructed the age ladder of \\cite{2007MNRAS.375.1220M} assigning new nominal ages, as shown in Table \\ref{ladder}. To enable the reader to add other SFRs to this ladder the pre-MS splines are freely available from the cluster collaboration home page\\footnote{http://www.astro.ex.ac.uk/people/timn/Catalogues/description.html}, and the CDS archive. \\item We have shown that metallicity information is now vital for accurate relative distances to SFRs. This is especially true if one is attempting to characterise evolutionary indicators such as disc fractions as a function of age, or trying to uncover environmental effects (such as the effect of ionising winds from massive stars on planet formation from discs). \\item We have discussed that the overlapping region of the R-C gap could be used to derive spreads in isochronal age in a CMD, i.e. the apparent age difference between the maximum mass star still on the convective pre-MS and the minimum mass star which has reached the MS. If star formation is slow and isochronal ages of individual stars are reliable this would provide a direct measurement of the age spreads present in SFRs. If star formation is rapid the R-C gap overlap region reveals the underlying spread in accretion histories within an SFR. This is important as for rapid star formation, if an accretion history is unknown isochronal ages derived from a position in a CMD do not represent the true age of a star. Indeed it is therefore likely that if a rapid star formation model is accurate median or mean ages drawn from a population are also invalid. A more useful approach may be to compare SFRs using age ladder arguments or perhaps to use the age of the youngest stars which have the lowest accretion history. \\item We have shown further evidence for non-uniform decay of discs in SFRs, although new comparisons must be made using consistent disc fraction indicators and mass ranges. \\end{enumerate}" }, "0801/0801.3346_arXiv.txt": { "abstract": "{% We combine extinction maps from the Two Micron All Sky Survey (2MASS) with Hipparcos and Tycho parallaxes to obtain reliable and high-precision estimates of the distance to the Ophiuchus and Lupus dark complexes. Our analysis, based on a rigorous maximum-likelihood approach, shows that the $\\rho$-Ophiuchi cloud is located at $(119 \\pm 6) \\mbox{ pc}$ and the Lupus complex is located at $(155 \\pm 8) \\mbox{ pc}$; in addition, we are able to put constraints on the thickness of the clouds and on their orientation on the sky (both these effects are not included in the error estimate quoted above). For Ophiuchus, we find some evidence that the streamers are closer to us than the core. The method applied in this paper is currently limited to nearby molecular clouds, but it will find many natural applications in the GAIA-era, when it will be possible to pin down the distance and three-dimensional structure of virtually every molecular cloud in the Galaxy. ", "introduction": "\\label{sec:introduction} In recent years, much effort has been dedicated to the study of dark molecular clouds and their dense cores. One of the main motivations for these investigations is the study of the process of star and planet formation in its entirety, and a deeper understanding of the effects of the local environment. A key aspect of the scientific analysis of a dark molecular cloud is its distance, which is related to many physically relevant properties (in particular, the mass scales as the square of the distance). Unfortunately, the distance estimates for many clouds are often plagued by very large uncertainties and it is not rare to see in the literature measurements that differ by large factors. Often, this quantity is inferred by associating the cloud to other astronomical objects whose distance is well known. For example, the Lupus complex is located near the Sco OB2 association \\citep{1978ApJS...38..309H, 1999AJ....117..354D}, whose distance is estimated to be $\\sim 150 \\mbox{ pc}$ based on the photometry of OB stars \\citep{1991ESOSR..11....1R}. However, this method is based on some degree of arbitrariness when making the link between the cloud and the other objects, and thus the deduced distance can be completely unreliable. The cloud complexes considered in this paper, the $\\rho$ Ophiuchi and the Lupus dark clouds, are good examples of how different distance estimates made by different authors can be. Quoted values for the $\\rho$-Ophiuchi cloud are in the range $120$--$165 \\mbox{ pc}$, as recently summarized by \\citet{2004AJ....127.1029R}. \\citet{1981A&A....99..346C} estimated the distance to be $\\sim 160 \\mbox{ pc}$ from multi-color photometry of heavily absorbed stars in the $\\rho$ Oph core, while \\citet{1998A&A...338..897K} suggested a significantly smaller figure, $\\sim 120 \\mbox{ pc}$. For Lupus, the situation is even more extreme, with estimates from $100 \\mbox{ pc}$ \\citep{1998A&A...338..897K} to $190 \\mbox{ pc}$ \\citep{1998MNRAS.301L..39W}, both based on Hipparcos data; the most widely accepted distance is in the middle of these two extremes, at approximately $140 \\mbox{ pc}$ \\citep{1993AJ....105..571H, 1999AJ....117..354D}. These uncertainties severely hamper our understanding of the physical properties of molecular clouds, and thus our knowledge of star formation. For example, an error by a factor two on the distance of a cloud translates into an error by a factor four in the mass, and has a thus a huge impact on the estimate of the density, size, stability, and star formation efficiency of their cloud cores (see \\citealp{2007A&A...462L..17A} for further discussions on this point). In this paper we present accurate distance measurements of the Ophiuchus and Lupus complexes based upon Hipparcos and Tycho parallaxes. Our technique is statistically sound and, when applied to nearby giant molecular clouds, is able to provide accurate distance measurements and related errors. The paper is organized as follows. In Sect.~\\ref{sec:basic-analysis} we present the main datasets used and discuss a simple approach to the problem. A more quantitative technique is developed in Sect.~\\ref{sec:likelihood-analysis}, where we also presents the results obtained for the two clouds considered in this paper. The technique is further developed in Sect.~\\ref{sec:second-order-geom} to include the effects of the cloud thickness and orientation. Finally, in Sect.~\\ref{sec:discussion} we briefly discuss the results obtained. ", "conclusions": "\\label{sec:discussion} Our analysis indicates that the Ophiuchus cloud is very likely to be closer than the we think: the ``standard'' distance, $160 \\mbox{ pc}$, is indeed excluded at $5 \\sigma$. Not surprisingly, our estimate is in excellent agreement with the one obtained by \\citet{1998A&A...338..897K} using a similar technique (but note that the two results are based on a different datasets and use different statistical methods). Recently, \\citet{WilsonPhD} and \\citet{2007arXiv0709.0505M} reported an Hipparcos estimated distance of Ophiuchus of $d = 136^{+8}_{-7} \\mbox{ pc}$ and $d = 135^{+8}_{-7} \\mbox{ pc}$, respectively, results which are marginally in agreement with our distance estimate of the Ophiuchus core. \\citeauthor{2007arXiv0709.0505M} mainly focused on the Lynds~168 cloud, and inferred the cloud distance from the parallaxes of seven stars presumably associated with illuminated nebulae. Although his method is plagued by a potential uncertainty, the real association between the Hipparcos stars and the Ophiuchus cloud complex, and suffers from the small number statistics (7 parallaxes), the agreement obtained is reassuring and suggests that we finally have in hand a reliable estimate of the Ophiuchus cloud complex. \\citet{WilsonPhD}, instead, used a more rigorous maximum-likelihood approach, but his analysis is substantially different from ours. In particular, his method is based on a preliminary classification of stars as foreground and background based on their observed reddening; the distance of the cloud is then inferred from the constraints imposed by the parallaxes of the stars considered. In contrast, in our method we do not need to explicitly classify stars, but on the contrary can use directly all data in the likelihood (an advantage of this is that we are able to use in our method also stars that cannot be uniquely classified as foreground or background because, e.g., of their large photometric errors). Very recently, \\citet{Loinard} used phase-referenced multi-epoch Very Long Baseline Array (VLBA) observations to measure the trigonometric parallax of the Ophiucus star-forming region. Their method is based on accurate (to better than a tenth of a milli-arcsecond), absolute measurements of the position of individual magnetically active young stars though to be associated with the star-forming region. If multi-epoch data are available, one can infer the parallax of the star, and thus of the star-forming regions, with exquisite accuracy (below $1 \\mbox{ pc}$ for objects within $\\sim 200 \\mbox{ pc}$). This technique was applied to four stars in the $\\rho$-Ophiuchi region, and interestingly it lead to somewhat contradictory results: two objects associated with the sub-condensation Oph~A (S1 and DoAr21) have a measured parallax close to $\\sim 120 \\mbox{ pc}$, and thus in excellent agreement with our estimate; two others, associated with the condensation Oph~B (VSSG14 and VL5), have a much higher distance of $\\sim 165 \\mbox{ pc}$. A possible explanation of this is provided by a closer examination of the two sources in the Oph~B condensation. From their spectral energy distributions (SEDs), they appear to be Class~III A-type stars, with little or no circumstellar emitting dust, and an extinction as large as $50 \\mbox{ mag}$ in the $V$ band (see \\citealp{1989ApJ...340..823W}, and especially \\citealp{1994ApJ...434..614G} for a discussion on the individual SEDs of VSSG14 and VL5). These results indicate that VSSG14 and VL5 are probably background stars, not directly associated with the Ophiucus complex, and likely to be part of a smaller OB association at $\\sim 165 \\mbox{ pc}$. The distance of the Lupus complex is substantially in agreement with the ones reported in the literature, but both the thickness analysis and the orientation analysis suggest noticeable differences in the various Lupus subclouds." }, "0801/0801.2389_arXiv.txt": { "abstract": " ", "introduction": "About one third of all known asteroids belong to families (\\cite{zappala}), which are clusters of asteroids believed to result from collisional disruption of parent bodies (\\cite{brien}). The hypothesis of collisional origin is based entirely on the observed similarity of dynamical and spectral properties, and not on the understanding of the collisional physics itself (\\cite{michel2001}). Members of asteroid families share similar orbital elements (semimajor axis, eccentricity and orbital inclination) (\\cite{nesvorny+}) and chemical composition (\\cite{juric}). They often have low density { which suggests a high porosity -- the asteroids are often made of} gravitationally bound fragments or ``rubble piles'' (\\cite{richardson}). { This structure is consistent with a collisional origin, and} explains both the lack of fast rotators among asteroids larger than a few hundred meters (\\cite{pravec}) and the presence of large craters revealed by close-up surface images (\\cite{chapman}). The rotational rates of the asteroids display a Maxwellian distribution, indicating angular momentum redistribution by collisions (\\cite{binzel,fulchignoni}). As such, prominent asteroid families are therefore important tracers of high-velocity impacts, one of the principal geologic processes affecting small bodies in the Solar System. Fig 1. comes here Collisions modify the shape and size of an asteroid. { The most energetic collisions disrupt the bodies and generate a number of smaller, gravitationally unbound fragments which will evolve further as new individual asteroids. The less energetic collisions can only fragment the body and/or form craters, without disruption. Collisions are rare, and therefore cannot be observed directly, and their complete understanding requires numerical experiments that have to reproduce the observable constraints of the impacts. Currently these constraints are: asteroid size distribution, the number of asteroid families, the age distribution of meteorites, and surface cratering of 5 asteroids which were encountered by spacecraft. The last of these cannot be easily interpreted, because craters can overlap or be completely saturated, and crater size depends on the impactor/crater scaling law model. The other constraints are only sensitive to the most energetic, disruptive collisions. Significant information on the less energetic impacts may be extracted from asteroid shapes: it has been proposed (\\cite{richardson,K+A}) that the shape of asteroids can evolve with age due to several microimpacts. However, there are only a few asteroids with known shape and hence the evolutionary effects could not have not been fully exploited yet.} While sizes can be constrained relatively easily from absolute photometry (\\cite{eddington}), determining shapes is much more difficult. Space missions have visited only a few asteroids and shapes for these have been determined accurately (\\cite{fujiwara,abe,okada,saito}). Radar observations have revealed shapes of Near-Earth Objects encountering the Earth ( e.g. \\cite{ostro}), while high-resolution imaging is restricted to the largest asteroids (\\cite{tanga,marchis}). For the rest of the asteroid belt, time-resolved photometry is the only method to determine the shape. The key effect is the rotation: it causes periodic brightness variations which, in principle, reveal the shape of a given asteroid if there are enough data for full inversion of the lightcurve, or at least for fitting a triaxial ellipsoid. Despite the relatively simple principles of the method, there has been slow progress because of the time-consuming nature of the observations. Recent advances in all-sky surveys offer new approaches to the problem. These surveys often observe thousands of asteroids in a night, and with the appropriate data processing the shape distribution can be constrained. Here we deduce the shape distribution of 11,735 main-belt asteroids belonging to eight prominent asteroid families, using data from the Sloan Digital Sky Survey (SDSS). SDSS was a deep imaging survey covering over a quarter of the Celestial Sphere in the Northen Galactic Cap, resulting in high-quality photometric measurements of 50 million stars and a similar number of galaxies. Although its main purpose was cosmology, SDSS has also collected 204,305 measurements of asteroids, listed in the SDSS Moving Object Catalog 3 (SDSS MOC) \\footnote{Available at http://www.sdss.org} (\\cite{ivezic}). Of these, 67,637 data points have been linked to 26,847 known asteroids (\\cite{juric}). This means that all the identified asteroids have been observed an average of 2.5 times. The calibration of the SDSS is very accurate because all measurements were acquired with the same instrument and from the same observing site. Therefore, systematic effects are minimized and the individual brightness measurements are accurate to 1--2\\%{} down to magnitude 20 in $r$ band. ", "conclusions": "" }, "0801/0801.3420_arXiv.txt": { "abstract": "The present standard model of cosmology states that the known particles carry only a tiny fraction of total mass and energy of the Universe. Rather, unknown dark matter and dark energy are the dominant contributions to the cosmic energy budget. We review the logic that leads to the postulated dark energy and present an alternative point of view, in which the puzzle may be solved by properly taking into account the influence of cosmic structures on global observables. We illustrate the effect of averaging on the measurement of the Hubble constant. ", "introduction": "The concordance model of cosmology states that only $4\\%$ of the mass/energy content of the Universe is carried by identified particles, and the dominant contributions are unknown dark matter ($22\\%$) and mysterious dark energy ($74\\%$) \\cite{Spergel}. It is said that we are living in the age of ``precision cosmology'', which is justified as a statement on the precision of observations, but certainly not on their understandings. The purpose of this talk is to recapitulate the genesis of the dark energy hypothesis and to show that it poses a serious crisis to theoretical cosmology. We argue that a possible resolution for this crisis may come from the proper treatment of effects from averaging. ", "conclusions": "" }, "0801/0801.1339_arXiv.txt": { "abstract": "A short discussion of theoretical results on cosmic ray first-order Fermi acceleration at relativistic shock waves is presented. We point out that the recent results by Niemiec with collaborators change the knowledge about these processes in a substantial way. In particular one can not expect such shocks to form particle distributions extending to very high energies. Instead, distributions with the shock compressed injected component followed by a more or less extended high energy tail are usually created. Increasing the shock Lorentz factor leads to steepening and decreasing of the energetic tail. Also, even if a given section of the spectrum preserves the power-law form, the fitted spectral index may be larger or smaller than the claimed `universal index' $\\sigma \\approx 2.2$~. A reported simple check of real shapes of electron spectra in the Cyg A hot spots provides results clearly deviating from the standard expectations for such shocks met in the literature. The spectrum consist of the very flat low energy part ($\\sigma \\approx 1.5$), up to electron energies $\\sim 1$ GeV, and much steeper part ($\\sigma > 3$) at higher energies. We conclude with presentation of a short qualitative discussion of the Fermi second-order processes acting in relativistic plasmas. We suggest that such processes can be the main accelerating agent for very high energy particles. In particular its can accelerate electrons to energies in the range of $1$ - $10^3$ TeV in relativistic jets, shocks and radio-source lobes. ", "introduction": "Relativistic plasma flows are observed in a number of astrophysical objects, ranging from a mildly relativistic jets of the sources like SS433, through the-Lorentz-factor-of-a-few jets in AGNs and galactic `micro-quasars', up to highly ultra-relativistic outflows in sources of gamma ray bursts or pulsar winds. As nearly all such objects are efficient emitters of nonthermal radiation, what requires existence of energetic particles, our attempts to understand the processes generating cosmic rays are essential for understanding the fascinating phenomena observed. Below I will discuss the work carried out in order to understand the cosmic ray acceleration processes acting at relativistic shocks and within highly turbulent regions accompanying such shocks and shear layers. I will not include here the interesting work involving collisionless shocks modelling with {\\it particle in cell} simulations. This approach uses quite different modelling method in comparison to the other work discussed here, relating in most cases to the characteristic energy range of shock compressed thermal plasma particles. The present paper is a modified and updated version of some of my previous reviews of the subject. Also, essentially the same slightly shortened text is provided as my contribution to the HEPRO Workshop in Dublin (September 2007). Below we will append the index `1' (`2') to quantities measured in the plasma rest frame upstream (downstream) of the shock. ", "conclusions": "" }, "0801/0801.4200_arXiv.txt": { "abstract": "Fluorescence light is induced by extensive air showers while developing in the Earth's atmosphere. The number of emitted fluorescence photons depends on the conditions of the air and on the energy deposited by the shower particles at every stage of the development. In a previous model calculation, the pressure and temperature dependences of the fluorescence yield have been studied on the basis of kinetic gas theory, assuming temperature-independent molecular collision cross-sections. In this work we investigate the importance of temperature-dependent collision cross-sections and of water vapour quenching on the expected fluorescence yield. The calculations will be applied to simulated air showers while using actual atmospheric profiles to estimate the influence on the reconstructed energy of extensive air showers. ", "introduction": "\\label{intro} Several air shower experiments like HiRes~\\cite{hires}, the Pierre Auger Observatory~\\cite{auger}, and Telescope Array~\\cite{ta}, are using the fluorescence technique for detecting extensive air showers (EAS) induced by ultra-high energy cosmic rays. Measuring the fluorescence light that nitrogen molecules emit after being excited by charged particles of EAS is currently the most direct method for determining the energy of EAS in a model-independent way. A thorough understanding of the light emission process is necessary to obtain the primary energy of EAS with high precision. In this paper, we extend our previous model calculation for the fluorescence light emission~\\cite{BK2006} by including the latest results on input parameters and their temperature dependence as obtained in laboratory measurements. For the reconstruction of air shower events, the light emission has to be known in dependence on altitude in the Earth's atmosphere at which the shower is observed. Up to now, the altitude dependence has been considered by including air density profiles and collisional quenching of nitrogen-nitrogen and nitrogen-oxygen molecules as described by kinetic gas theory. The cross-sections for collisional quenching were taken to be temperature independent. However, the cross-sections describing collisional quenching are known to be temperature-dependent~\\cite{gs}. Gr\\\"un and Schopper~\\cite{gs} found a decreasing collisional quenching cross-section with increasing temperature. Recently, the AirFly experiment has studied collisional quenching cross-sections in dependence on temperature~\\cite{airfly_icrc}. These data are also included in the model calculations presented in this article. In addition we investigate the influence of water vapour on the fluorescence yield, using relative humidity measurements performed at the site of the Auger detector in Argentina. ", "conclusions": "\\label{concl} The effects of temperature-dependent collisional quenching cross-sections and of quenching due to water vapour have been studied. Both effects lead to a significant reduction of the fluorescence yield in the lower part of the atmosphere. Applying these calculations to simulated EAS, a distortion of the longitudinal shower development is found. A reduction of the emitted ligth is expected, which varies from about 7\\% to 11\\% depending on seasonal atmospheric model and on zenith angle of the EAS. Hence, accounting for these effects in the reconstruction of the primary energy of EAS, the primary energy will be increased by this amount as compared with the former model calculations. The position of the shower maximum is hardly shifted, in all atmospheric models the shift is less than 50~m. \\subsection*{Acknowledgement} On of the authors (BK) is supported by the German Research Foundation (DFG) under contract KE 1151/1-2." }, "0801/0801.4493_arXiv.txt": { "abstract": "APEX single-dish observations at sub-millimeter wavelengths toward a sample of massive star-forming regions reveal that C$_2$H is almost omni-present toward all covered evolutionary stages from Infrared Dark Clouds via High-Mass Protostellar Objects to Ultracompact H{\\sc ii} regions. High-resolution data from the Submillimeter Array toward one hot-core like High-Mass Protostellar Object show a shell-like distribution of C$_2$H with a radius of $\\sim$9000\\,AU around the central submm peak position. These observed features are well reproduced by a 1D cloud model with power-law density and temperature distributions and a gas-grain chemical network. The reactive C$_2$H radical (ethynyl) is abundant from the onset of massive star formation, but later it is rapidly transformed to other molecules in the core center. In the outer cloud regions the abundance of C$_2$H remains high due to constant replenishment of elemental carbon from CO being dissociated by the interstellar UV photons. We suggest that C$_2$H may be a molecule well suited to study the initial conditions of massive star formation. ", "introduction": "Spectral line surveys have revealed that high-mass star-forming regions are rich reservoirs of molecules from simple diatomic species to complex and larger molecules (e.g., \\citealt{schilke1997b,hatchell1998b,comito2005,bisschop2007}). However, there have been rarely studies undertaken to investigate the chemical evolution during massive star formation from the earliest evolutionary stages, i.e., from High-Mass Starless Cores (HMSCs) and High-Mass Cores with embedded low- to intermediate-mass protostars destined to become massive stars, via High-Mass Protostellar Objects (HMPOs) to the final stars that are able to produce Ultracompact H{\\sc ii} regions (UCH{\\sc ii}s, see \\citealt{beuther2006b} for a recent description of the evolutionary sequence). The first two evolutionary stages are found within so-called Infrared Dark Clouds (IRDCs). While for low-mass stars the chemical evolution from early molecular freeze-out to more evolved protostellar cores is well studied (e.g., \\citealt{bergin1997,dutrey1997,pavlyuchenkov2006,joergensen2007}), it is far from clear whether similar evolutionary patterns are present during massive star formation. To better understand the chemical evolution of high-mass star-forming regions we initiated a program to investigate the chemical properties from IRDCs to UCH{\\sc ii}s from an observational and theoretical perspective. We start with single-dish line surveys toward a large sample obtaining their basic characteristics, and then perform detailed studies of selected sources using interferometers on smaller scales. These observations are accompanied by theoretical modeling of the chemical processes. Long-term goals are the chemical characterization of the evolutionary sequence in massive star formation, the development of chemical clocks, and the identification of molecules as astrophysical tools to study the physical processes during different evolutionary stages. Here, we present an initial study of the reactive radical ethynyl (C$_2$H) combining single-dish and interferometer observations with chemical modeling. Although C$_2$H was previously observed in low-mass cores and Photon Dominated Regions (e.g., \\citealt{millar1984,jansen1995}), so far it was not systematically investigated in the framework of high-mass star formation. ", "conclusions": "To understand the observations, we conducted a simple chemical modeling of massive star-forming regions. A 1D cloud model with a mass of 1200\\,M$_\\sun$, an outer radius of 0.36\\,pc and a power-law density profile ($\\rho\\propto r^p$ with $p=-1.5$) is the initially assumed configuration. Three cases are studied: (1) a cold isothermal cloud with $T=10$\\,K, (2) $T=50$\\,K, and (3) a warm model with a temperature profile $T\\propto r^q$ with $q=-0.4$ and a temperature at the outer radius of 44\\,K. The cloud is illuminated by the interstellar UV radiation field (IRSF, \\citealt{draine1978}) and by cosmic ray particles (CRP). The ISRF attenuation by single-sized $0.1\\mu$m silicate grains at a given radius is calculated in a plane-parallel geometry following \\citet{vandishoeck1988}. The CRP ionization rate is assumed to be $1.3\\times 10^{-17}$~s$^{-1}$ \\citep{spitzer1968}. The gas-grain chemical model by \\citet{vasyunin2008} with the desorption energies and surface reactions from \\citet{garrod2006} is used. Gas-phase reaction rates are taken from RATE\\,06 \\citep{woodall2007}, initial abundances, were adopted from the ``low metal'' set of \\citet{lee1998}. Figure \\ref{model} presents the C$_2$H abundances for the three models at two different time steps: (a) 100\\,yr, and (b) in a more evolved stage after $5\\times10^4$\\,yr. The C$_2$H abundance is high toward the core center right from the beginning of the evolution, similar to previous models (e.g., \\citealt{millar1985,herbst1986,turner1999}). During the evolution, the C$_2$H abundance stays approximately constant at the outer core edges, whereas it decreases by more than three orders of magnitude in the center, except for the cold $T=10$~K model. The C$_2$H abundance profiles for all three models show similar behavior. The chemical evolution of ethynyl is determined by relative removal rates of carbon and oxygen atoms or ions into molecules like CO, OH, H$_2$O. Light ionized hydrocarbons CH$^+_{\\rm n}$ (n=2..5) are quickly formed by radiative association of C$^+$ with H$_2$ and hydrogen addition reactions: C$^+$ $\\rightarrow$ CH$_2^+$ $\\rightarrow$ CH$_3^+$ $\\rightarrow$ CH$_5^+$. The protonated methane reacts with electrons, CO, C, OH, and more complex species at later stage and forms methane. The CH$_4$ molecules undergo reactive collisions with C$^+$, producing C$_2$H$_2^+$ and C$_2$H$_3^+$. An alternative way to produce C$_2$H$_2^+$ is the dissociative recombination of CH$_5^+$ into CH$_3$ followed by reactions with C$^+$. Finally, C$_2$H$_2^+$ and C$_2$H$_3^+$ dissociatively recombine into CH, C$_2$H, and C$_2$H$_2$. The major removal for C$_2$H is either the direct neutral-neutral reaction with O that forms CO, or the same reaction but with heavier carbon chain ions that are formed from C$_2$H by subsequent insertion of carbon. At later times, depletion and gas-phase reactions with more complex species may enter into this cycle. At the cloud edge the interstellar UV radiation instantaneously dissociates CO despite its self-shielding, re-enriching the gas with elemental carbon. The transformation of C$_2$H into CO and other species proceeds efficiently in dense regions, in particular in the ``warm'' model where endothermic reactions result in rich molecular complexity of the gas (see Fig.~\\ref{model}). In contrast, in the ``cold'' 10\\,K model gas-grain interactions and surface reactions become important. As a result, a large fraction of oxygen is locked in water ice that is hard to desorb ($E_{\\rm des} \\sim 5500$~K), while half of the elemental carbon goes to volatile methane ice ($E_{\\rm des} \\sim 1300$~K). Upon CRP heating of dust grains, this leads to much higher gas-phase abundance of C$_2$H in the cloud core for the cold model compared to the warm model. The effect is not that strong for less dense regions at larger radii from the center. Since the C$_2$H emission is anti-correlated with the dust continuum emission in the case of IRAS\\,18089-1732 (Fig.\\,\\ref{18089}), we do not have the H$_2$ column densities to quantitatively compare the abundance profiles of IRAS\\,18089-1732 with our model. However, data and model allow a qualitative comparison of the spatial structures. Estimating an exact evolutionary time for IRAS\\,18089-1732 is hardly possible, but based on the strong molecular line emission, its high central gas temperatures and the observed outflow-disk system \\citep{beuther2004a,beuther2004b,beuther2005c}, an approximate age of $5\\times10^4$\\,yr appears reasonable. Although dynamical and chemical times are not necessarily exactly the same, in high-mass star formation they should not differ to much: Following the models by \\citet{mckee2003} or \\citet{krumholz2006b}, the luminosity rises strongly right from the onset of collapse which can be considered as a starting point for the chemical evolution. At the same time disks and outflows evolve, which should hence have similar time-scales. The diameter of the shell-like C$_2$H structure in IRAS\\,18089-1732 is $\\sim 5''$ (Fig.\\,\\ref{18089}), or $\\sim$9000\\,AU in radius at the given distance of 3.6\\,kpc. This value is well matched by the modeled region with decreased C$_2$H abundance (Fig.\\,\\ref{model}). Although in principle optical depths and/or excitation effects could mimic the C$_2$H morphology, we consider this as unlikely because the other observed molecules with many different transitions all peak toward the central submm continuum emission in IRAS\\,18089-1732 \\citep{beuther2005c}. Since C$_2$H is the only exception in that rich dataset, chemical effects appear the more plausible explanation. The fact that we see C$_2$H at the earliest and the later evolutionary stages can be explained by the reactive nature of C$_2$H: it is produced quickly early on and gets replenished at the core edges by the UV photodissociation of CO. The inner ``chemical'' hole observed toward IRAS\\,18089-1732 can be explained by C$_2$H being consumed in the chemical network forming CO and more complex molecules like larger carbon-hydrogen complexes and/or depletion. The data show that C$_2$H is not suited to investigate the central gas cores in more evolved sources, however, our analysis indicates that C$_2$H may be a suitable tracer of the earliest stages of (massive) star formation, like N$_2$H$^+$ or NH$_3$ (e.g., \\citealt{bergin2002,tafalla2004,beuther2005a,pillai2006}). While a spatial analysis of the line emission will give insights into the kinematics of the gas and also the evolutionary stage from chemical models, multiple C$_2$H lines will even allow a temperature characterization. With its lowest $J=1-0$ transitions around 87\\,GHz, C$_2$H has easily accessible spectral lines in several bands between the 3\\,mm and 850\\,$\\mu$m. Furthermore, even the 349\\,GHz lines presented here have still relatively low upper level excitation energies ($E_u/k\\sim42$\\,K), hence allowing to study cold cores even at sub-millimeter wavelengths. This prediction can further be proved via high spectral and spatial resolution observations of different C$_2$H lines toward young IRDCs." }, "0801/0801.1080_arXiv.txt": { "abstract": "The line-of-sight velocities and [OIII] 5007 \\AA\\ expansion velocities are measured for 11 planetary nebulae (PNs) in the Virgo cluster core, at 15 Mpc distance, with the FLAMES spectrograph on the ESO VLT. These PNs are located about halfway between the two giant ellipticals M87 and M86. From the [OIII] 5007 \\AA\\ line profile widths, the average half-width at half maximum expansion velocity for this sample of 11 PNs is $ \\bar{v}_{HWHM} = 16.5$ kms$^{-1}$ (RMS = 2.6 kms$^{-1}$). We use the PN subsample bound to M87 to remove the distance uncertainties, and the resulting [OIII] 5007 \\AA\\ luminosities to derive the central star masses. We find these masses to be at least $0.6$ M$_\\odot$ and obtain PN observable life times $t_{PN} < 2000 $ yrs, which imply that the bright PNs detected in the Virgo cluster core are compact, high density nebulae. We finally discuss several scenarios for explaining the high central star masses in these bright M87 halo PNs. ", "introduction": "Since the discovery of free-floating intracluster planetary nebulae (ICPNs) in the Virgo cluster \\citep{Arnaboldi+96}, extensive imaging and spectroscopic observations were carried out to determine their projected phase space distribution and the fraction of diffuse cluster light not bound to Virgo galaxies. To enlarge the sample of more than 40 ICPN line-of-sight (LOS) velocities available from \\citet{Arnaboldi+03,Arnaboldi+04}, we have obtained new PN spectra with FLAMES at the ESO VLT \\citep{Doherty+08}. The high spectral resolution of the new data allows us for the first time to measure expansion velocities for the [OIII] nebula shells of 11 PNs in the Virgo cluster core, nearly half way between M87 and M86. The expansion velocity of a planetary nebula (PN) is one of the most important parameters determining its evolution, but currently it is known only for a few hundred Galactic PNs, mostly bright objects in the Milky Way Disk \\citep{gesickizijlstra00}, and for a few tens in the Magellanic Clouds and M31. When interpreted using dynamically evolving nebular models \\citep[e.g.][]{Schoenberner+05}, PN shell expansion velocities provide reliable extimates of PN dynamical ages and distance estimates for Galactic PNs. In this letter, we give a brief summary of our Observations in Section~\\ref{obser}. In Section~\\ref{PNexp}, we present the PN [OIII] 5007 \\AA\\ half-width half maximum velocity measurements $v_{HWHM}$, and $m(5007)$ magnitudes as defined by \\citet{Jacoby89}. In Section~\\ref{M87PNLF} we build the PN luminosity function (PNLF) for the spectroscopically confirmed PN sample in Virgo and for the subsample bound to the M87 halo. In Section~\\ref{outerradius}, we then estimate the outer radii of the M87 halo PNs from the observed PN expansion velocities and their visibility lifetimes. We finally discuss in Section~\\ref{fdiscuss} the remaining distance uncertainties, the shape of the PNLF in the M87 halo, and the mechanisms that may be responsible for the large core masses of the brightest PNs in the halo of M87. ", "conclusions": "We have measured the nebular [OIII] 5007\\AA\\ spectroscopic expansion velocities for 11 PNs in the Virgo cluster core, at 15 Mpc distance, with the FLAMES spectrograph on the ESO VLT. Based on the [OIII] line profile width, the average spectroscopic expansion velocity for this sample is $ \\bar{v}_{HWHM} = 16.5$ kms$^{-1}$ (RMS = 2.6 kms$^{-1}$), which is in agreement with the predictions of dynamically evolving nebular models for high density nebulae close to their maximal $m(5007)$ emission. Large central star masses M$_{CS} > 0.6$~M$_\\odot$ are inferred from the bright measured [OIII] luminosities and the known distance for the PNs bound to the M87 halo. From the large central star masses and the measured $v_{HWHM}$, we derive short PN visibility times and small nebular outer radii, $ \\sim 0.07$ pc. The PNs in the M87 halo have large central star masses, are compact and their nebula shells may be similar to those observed for the brightest Galactic Bulge PNs. Three mechanisms are reviewed as possible explanation for the large core masses of the M87 halo PNs: intermediate age population, metallicity effects, and blue stragglers, with the latter being the most likely, given the old age and the low metallicities of the IRGB stars in the Virgo core \\citep{Williams+07}." }, "0801/0801.1755_arXiv.txt": { "abstract": "We study the dynamical behaviour of the interacting holographic dark energy model whose interaction term is $Q=3H(\\lam_d\\rho_d + \\lam_c\\rho_c)$, where $\\rho_d$ and $\\rho_c$ are the energy density of dark energy and CDM respectively. To satisfy the observational constraints from SNIa, CMB shift parameter and BAO measurement, if $\\lam_c = \\lam_d$ or $\\lam_d, \\lam_c >0$, the cosmic evolution will only reach the attractor in the future and the ratio $\\rho_c/\\rho_d$ cannot be slowly varying at present. Since the cosmic attractor can be reached in the future even when the present values of the cosmological parameters do not satisfy the observational constraints, the coincidence problem is not really alleviated in this case. However, if $\\lam_c \\neq \\lam_d$ and they are allowed to be negative, the ratio $\\rho_c/\\rho_d$ can be slowly varying at present and the cosmic attractor can be reached near the present epoch. Hence, the alleviation of the coincidence problem is attainable in this case. The alleviation of coincidence problem in this case is still attainable when confronting this model to SDSS data. \\vspace{3mm} \\begin{flushleft} \\textbf{Keywords}: Dark Energy. \\end{flushleft} ", "introduction": "Observations suggest that the expansion of the universe is accelerating \\cite{Riess:98,Perl:99}. The acceleration of the universe may be explained by supposing that the present universe is dominated by a mysterious form of energy whose pressure is negative, known as dark energy. One problem of the dark energy model is the coincidence problem, which is the problem why the dark energy density and matter density are of the same order of magnitude in the present epoch although they differently evolve during the expansion of the universe. A possible way to alleviate the coincidence problem is to suppose that there is an interaction between matter and dark energy. The cosmic coincidence can then be alleviated by appropriate choice of the form of the interaction between matter and dark energy leading to a nearly constant ratio $r=\\rho_c/\\rho_d$ during the present epoch \\cite{Zimdahl:03, Campo:06, Sadjadi:06} or giving rise to attractor of the cosmic evolution at late time \\cite{Amen:99, Chimento:03}. Since the existence of the cosmic attractor implies constant $r$ but the attractor does not always occur at the present epoch, we first find a range of dark energy parameters for which the attractor occurs and then check the evolution of $r$ during the present epoch. Based on holographic ideas \\cite{Cohen:98, Li:04}, one can determine the dark energy density in terms of the horizon radius of the universe. This type of dark energy is holographic dark energy \\cite{Setare:07} - \\cite{Elizalde:05}. By choosing Hubble radius as the cosmological horizon, the present amount of dark energy density agrees with observations. Nevertheless, dark energy evolves like matter at present, so it cannot lead to an accelerated expansion. However, if the particle horizon is chosen to be the cosmological horizon, the equation of state parameter of dark energy can become negative but not negative enough to drive an accelerating universe. The situation is better when one uses the event horizon as the cosmological horizon. In this case, dark energy can drive the present accelerated expansion, and the coincidence problem can be resolved by assuming an appropriate number of e-foldings of inflation. Roughly speaking, the coincidence problem can be resolved because the size of the cosmological horizon during the present epoch depends on the amount of e-folds of inflation, and the amount of holographic dark energy depends on the horizon size. Nevertheless, the second law of thermodynamics will be violated if $w_d < -1$ \\cite{Li:04, Gong:06}. Hence, $w_d$ should not cross the boundary $w_d = -1$. The boundary $w_d = -1$ can be crossed if dark energy interacts with matter. Since now the horizon size has a dependence on the interaction terms,the alleviation of cosmic coincidence should also depend on the interaction term. In this work we suppose that the holographic dark energy interact only with cold dark matter (CDM) and treat baryons as non-interacting matter component. Our objective is to compare the region of dark energy parameters for which the cosmic evolution has an attractor within the parameter region that satifies the observational constraints from combined analysis of SNIa data \\cite{Riess:06}, CMB shift parameter \\cite{Wang:06} and BAO measurement \\cite{Eisenstein:05}. The results of the comparison can tell us about the range of parameters that alleviate the cosmic coincidence. ", "conclusions": "For the interacting holographic dark energy model, we study the fixed points and their stabiliby, and compare a range of model parameters for which attractor exists with the $99.7\\%$ confidence levels from the combined analysis of SNIa data, CMB shift parameter and BAO measurement. Neglecting baryons, the observational constraints require that the value of $\\Om_c$ at the attractor point must be small if $\\lam_d = \\lam_c$ or $\\lam_d, \\lam_c >0$. This implies that the cosmic evolution will reach the attractor point in the future when $\\Om_c$ becomes small. In this case, $r$ cannot be slowly varying during the present epoch and the cosmic attractor cannot be reached near the present. Hence, the coincidence problem is not really alleviated for this case. However, if $\\lam_d$ and $\\lam_c$ are allowed to be negative, the cosmic evolution can reach the attractor near the present epoch for a narrow range of $\\lam_d$ and $\\lam_c$. Therefore, the coincidence problem is possible to alleviate in this case. Including baryons in our consideration, the attractor of the cosmic evolution cannot occur at present due to the non-vanishing baryon fraction. According to observations, the fixed point in this case is possible only when $Q$ is small and positive. Hence, the fixed point will be slowly reached in the future. These results indicate that for the interacting holographic dark energy model with the interaction terms considered here, the cosmic coincidence problem cannot be alleviated very well. We also briefly considered the constraint from SDSS matter power spectrum on the dark energy parameters. We have found that the parameters ranges that lead to the alleviation of cosmic coincidence are allowed by SDSS data." }, "0801/0801.0279_arXiv.txt": { "abstract": "Polarimetry is extensively used as a tool to trace the interstellar magnetic field projected on the plane of sky. Moreover, it is also possible to estimate the magnetic field intensity from polarimetric maps based on the Chandrasekhar-Fermi method. In this work, we present results for turbulent, isothermal, 3-D simulations of sub/supersonic and sub/super-Alfvenic cases. With the cubes, assuming perfect grain alignment, we created synthetic polarimetric maps for different orientations of the mean magnetic field with respect to the line of sight (LOS). We show that the dispersion of the polarization angle depends on the angle of the mean magnetic field regarding the LOS and on the Alfvenic Mach number. However, the second order structure function of the polarization angle follows the relation $SF \\propto l^{\\alpha}$, $\\alpha$ being dependent exclusively on the Alfvenic Mach number. The results show an anti-correlation between the polarization degree and the column density, with exponent $\\gamma \\sim -0.5$, in agreement with observations, which is explained by the increase in the dispersion of the polarization angle along the LOS within denser regions. However, this effect was observed exclusively on supersonic, but sub-Alfvenic, simulations. For the super-Alfvenic, and the subsonic model, the polarization degree showed to be intependent on the column density. Our major quantitative result is a generalized equation for the CF method, which allowed us to determine the magnetic field strength from the polarization maps with errors $< 20\\%$. We also account for the role of observational resolution on the CF method. ", "introduction": "It is believed that giant molecular clouds in the interstellar medium (ISM) are threaded by large scale magnetic fields \\citep{sch98, cru99}. However, it is still not completely clear what is the role of the magnetic field in the dynamics of the ISM and what is its effect on the star formation process. Also, the ratio of the magnetic and turbulent energy in these environments is a subject of controversy \\citep{padoan02, girart06}. Magnetic fields can influence the injection and evolution of turbulence bringing more complexity to this issue (see Lazarian \\& Cho [2004] for review). As an example, simulations have shown that strongly magnetized turbulent media develop structures with lower density contrasts when compared to pure hydrodynamic turbulence \\citep{kowal07, kritsuk07}. Observationally, different techniques can be used to measure the ISM magnetic field and determine its intensity and topology. Zeeman splitting of spectral lines provides a direct and precise derivation of the magnetic field component along the line of sight (LOS), mainly for clouds presenting strong spectral lines \\citep{heiles05}. However, it cannot be applied to the clouds where the line intensities are too weak. Typically, when observed, Zeeman measurements of molecular clouds give $B_{\\rm LOS} \\sim 10^{1-3} \\mu$G, and suggest the correlation with density $B_{\\rm LOS} \\propto \\rho^{0.5}$, which is consistent to the expected relation for compressions of magnetic fields frozen into plasma. Spectral line broadening show that molecular clouds present supersonic, but critically Alfvenic motions \\citep{cru99b}. This fact shows that the turbulent motions may be excited by MHD modes instead of being purely hydrodynamical. One of the most readily available methods of studying the perpendicular component of the magnetic field is based on the polarization of dust thermal emissions at infrared and submillimetric wavelengths \\citep{hil00}. The alignment of grains in respect to the magnetic field is a hot research topic (see Lazarian [2007] for review). Radiative torques (RATs) can promote alignment of irregular dust particles, resulting in different intensities for polarized radiation parallel and perpendicular to the local magnetic field \\citep{dolg76, draine96, draine97, lazho07a}. Grains with long axis aligned perpendicular to the magnetic field induce polarization parallel to the magnetic field for transmitted star light, and perpendicular to the field lines for the dust emission. \\cite{cho05} showed that RATs are very efficient on the grain alignment process in molecular clouds, even for the very dense regions (up to $A_V<10$). They also showed that the alignment efficiency strongly depends on the grain size, being practically perfect for large grains ($a>0.1$ $\\mu$m). More detailed studies of the RATs efficiency by \\cite{lazho07a} confirmed this claim. Therefore, for a range of $A_V$ it is acceptable to assume that the grains are well-aligned. For a given polarization map of an observed region, the mean polarization angle indicates the orientation of the large scale magnetic field. On the other hand the polarization dispersion gives clues on the value of the turbulent energy. This, as a consequence, can be used to determine the magnetic field component along the plane of sky. \\cite{chandra53} introduced a method (CF method hereafter) for estimating the ISM magnetic fields based on the dispersions of the polarization angle and gas velocity. Simply, it is assumed that the magnetic field perturbations are Alfvenic and that the rms velocity is isotropic. A promising approach to test this method is to create two-dimensional (plane of sky) synthetic maps from numerically simulated cubes. \\cite{ostriker01} performed 3D-MHD simulations, with $256^3$ resolution, in order to obtain polarization maps and study the validity of the CF method on the estimation of the magnetic field component along the plane of sky. They showed that the CF method gives reasonable results for highly magnetized media, in which the dispersion of the polarization angle is $< 25^{\\circ}$. However, they did not present any other statistical analysis or predictions that could be useful for the determination of the ISM magnetic field from observations. \\cite{heitsch01} presented a complementary work, with a more detailed analysis regarding the limited observational resolution on the CF method, and presented a modified equation to account for the differences obtained previously. They concluded that lower observational resolution leads to an overestimation of the magnetic field from the CF equation. They also showed a good agreement between the CF technique and the expected magnetic field of their simulations, except for the weak field models. Polarization maps from numerical simulations can also be used in the study of the correlation between the polarization degree and the total emission intensity (or dust column density). Observationally, the polarization degree in dense molecular clouds decreases with the total intensity as $P \\propto I^{-\\alpha}$, with $\\alpha = 0.5 - 1.2$ \\citep{goncalves05}. \\cite{padoan01} studied the role of turbulent cells in the $P \\ versus \\ I$ relation using supersonic and super-Alfvenic self-graviting MHD simulations. They found a decrease of polarization degree with total dust emission within gravitational cores, in agreement with observations, if grains are assumed to be unaligned for $A_V > 3$. When the alignment was assumed to be independent on $A_V$, the anti-correlation was not observed. Recently, \\cite{pelkonen07} extended this work and refined the calculation of polarization degree introducing the radiative transfer properly. In that work, the decrease in the alignment efficiency arises without any {\\it ad hoc} assumption. The alignment efficiency decreases as the radiative torques become less important in the denser regions. However, it is still not clear the role of the magnetic field topology and the presence of multiple cores intercepted by the line of sight on the decrease of polarization degree. In this work we attempt to extend the previously cited studies improving and applying the CF method for different situations. For that, we studied both sub and super-Alfvenic models, to study the role of the magnetic field topology in the observed polarization maps. We simulate different observational resolutions in the calculations of polarization maps and provide combined statistical analysis for both dust absorption and emission maps. We also present statistics based methods to characterize the turbulence and magnetic properties from polarization maps. We performed numerical simulations of magnetized turbulent plasma with higher resolution, which are described in Sec.\\ 2. From the data, we computed ``observable\" polarization maps, as shown in Sec.\\ 3. We then present the statistics and spatial distributions of angle and polarization degree for different models in Sec.\\ 4. In Sec.\\ 5 we propose the generalized equation for the CF method and compare it with the expected values to study its validity. In Sec.\\ 6, we discuss the improved procedures of polarization vector statistics, which allow observers to characterize the mean and fluctuating magnetic field of the cloud. We also discuss the applicability of our approach for polarized molecular and atomic lines, and compare the results with previous works. Ou summary is provided in Sec.\\ 7. ", "conclusions": "Emission and extinction polarimetric measurements provide an unique technique for the study of the magnetic field, projected into the plane of sky, in molecular clouds. Synthetic extinction maps depend on additional assumptions about the stellar population and may be more explored in a future work. In this work we focused on providing synthetic emission polarimetric maps, as well as different statistical analysis that could be used in the future by observers to infer the physical properties of the studied region. The physical interpretations of our results, as well as the comparisons with previous theoretical works, are given as follows. \\subsection{Our models} In this work we presented four different models: (1) $\\beta = 1.0$, $M_{\\rm S}=0.7$ and $M_{\\rm A}=0.7$, (2) $\\beta = 0.1$, $M_{\\rm S}=2.0$ and $M_{\\rm A}=0.7$, (3) $\\beta = 0.01$, $M_{\\rm S}=7.0$ and $M_{\\rm A}=0.7$ and (4) $\\beta = 0.1$, $M_{\\rm S}=7.0$ and $M_{\\rm A}=2.0$. Similar studies provided by \\cite{ostriker01} characterized different models by their pressure ratio. However, our results show completely different polarization maps for the two coincident $\\beta$-value models. This because the super-Alfvenic flows tend to tangle the magnetic field lines, what is not seen in the sub-Alfvenic models (Model 2), even with similar pressure ratio. The super-Alfvenic case shows a randomly distributed column density maps, with high constrast between the denser and rarefied regions. On the other hand, sub-Alfvenic cases are more filamentary, with contrasts increasing with the sonic Mach number. This general picture is independent on the angle between the external mean magnetic field and the LOS $\\theta$. However, for $B_{ext}$ nearly parallel to the LOS, the observed polarization will mostly be due to the random fluctuation component $\\delta B$. This effect is noticeable comparing Figs. 1 and 2. We see that for sub-Alfvenic turbulence the large scale density enhancements are mostly parallel to the mean magnetic fields, with exception to the very dense cores, which can easily change the orientation of the magnetic field. As a consequence, polarization maps will present dense structures mostly aligned with the mean magnetic field. This effect also play a role on the generation of the polarization maps. Since we integrate the polarization vectors along the LOS, the low density cells will systematically increase the homogeneous contribution, as well as the resulting polarization degree. To avoid this effect, we disregarded the contribution from low density cells using a threshold, which depends on the model used. For the models where the magnetic field is oriented parallel to the LOS, the polarization maps will show polarization vectors randomly oriented in respect to the density structures. It reveals the degeneracy on the polarimetric maps between the super-Alfvenic models with those with $B$ nearly parallel to the LOS. \\subsection{Polarization degree versus emission intensity} The polarization maps showed that the polarization degree is anti-correlated to the column density, in exception to the subsonic case. This result is in agreement with the observations, which revealed ``polarization holes\" associated to the dense cores for most of the regions observed. Observationally, \\cite{wolf03} showed that the polarization maps of molecular clouds follow the relation $P \\propto I^{-\\gamma}$, with $\\gamma \\sim 0.5 - 1.2$. The same trend is observed from polarized extinction of background stars \\citep{arce98}. It was proposed that this behavior occurs due to changes on dust properties inside denser cores, or even by an increase in thermal pressure, causing depolarization. \\cite{cho05} studied the role of the radiative torques on the grain alignment at dense cores and obtained $\\gamma \\sim 0.5 - 1.5$, depending on the dust size distribution. \\cite{padoan01} obtained a similar behavior from their numerical simulations of protostellar cores, though for only three dense cores of one single simulation. They assumed a cut-off on grain alignment efficiency for $A_V > 3$mag. For their model, if the alignment efficiency is independent on $A_V$, the polarization degree was shown to be independent on the column density. \\cite{pelkonen07} extended this work, improving the radiative transfer. They naturally obtained a decrease in the grain alignment at denser regions, explaining the lower degree of polarization. In our models, we assumed perfect grain alignment (independent on $A_V$). Therefore, the depolarization is exclusively due to the dispersion increase of the polarization angles in denser regions. We obtained $\\gamma \\sim 0.5$ for Models (2) and (3), but no correlation (i.e.\\ $\\gamma \\sim 0$) was found for Models (1) and (4). For Model (1), even the densest cores are unable to tangle the magnetic field lines and the polarization degree is homogeneously large. For Model (4), we have the opposite situation. The super-Alfvenic turbulence causes a strong dispersion of the magnetic field even at the less dense regions. For this case, the polarization degree is low everywhere. For Models (2) and (3), the turbulence is unable to destroy the magnetic field structure, but is able to create the dense cores by shocks. The cores are dense enough to drag the magnetic field lines and to increase the local $M_A$. Possibly, our results differ from the obtained by \\cite{padoan01} because of: i - numerical resolution, ii - $M_A$, and iii - self-gravity. We used $512^3$ simulations (instead of a $128^3$) and, as a result, our magnetic field structure is less homogeneous and the density constrast is higher. The larger complexities present in our cubes increase the effect described in the previous paragraph. \\cite{padoan01} and \\cite{pelkonen07} used a single, super-Alfvenic, model. We showed that the polarization maps for $M_A$ present a flat $P \\times I$ correlation. Finally, self-gravity causes the colapse of the denser regions compressing the magnetic field within these cores. As a consequence, if no strong diffusion takes place, the polarization degree tends to grow. As a future work, we plan to study properly the depolarization at dense cores considering the grain alignment process and self-gravity. \\subsection{Statistics of polarization angles} We found that the distributions of polarization angles of sub-Alfvenic models are similar, even for different magnetic to gas pressure ratios. However, the dispersion of angles increases with $M_{\\rm A}$ and with the inclination of the external magnetic field regarding the line of sight ($\\theta$). Actually, we noticed that the critical parameter is the Alfvenic Mach number considering the magnetic field component projected into the plane of sky (i.e. $M_{\\rm A}^{\\rm sky} = \\delta v \\sqrt{4 \\pi \\rho}/B_{\\rm sky}$). We can compare these results with \\cite{padoan01} and \\cite{ostriker01}. The first used one single super-Alfvenic model, and obtained irregular (flat) distributions of polarization angle. The latest analysed supersonic models for $\\beta = 0.01, 0.1$ and 1.0. They obtained clearly gaussian distributions for $\\beta = 0.01$, with increasing dispersion for larger inclinations. Also, they obtained flatter distributions as $\\beta$ increases (i.e. as the Alfvenic Mach number increases). These are all in agreement with our results. On the other hand, the power spectra analysis was showed to depend on the sonic Mach number. The spectra of the polarization angles show an increase in the power of small scales for increasing $M_{\\rm S}$. This occurs due to the amplification on the perturbations of the smallest scales for stronger turbulence. However, the same behavior is seen varying the inclination of the mean magnetic field. In this sense, there is a degeneracy between the Alfvenic Mach number and the orientation of the magnetic field. In this sense, structure functions of the polarization angle showed to be useful to avoid this degeneracy. The sub-Alfvenic models presented SFs with slope $\\alpha \\sim 0.5$ ($SF \\propto l^{\\alpha}$), independent on the magnetic field orientation. On the other hand, the super-Alfvenic model presented flatter SFs ($\\alpha \\sim 0.3$). Therefore, we could conjecture that it could be possible to obtain the mean magnetic field intensity, independently on its orientation regarding the LOS, from the SF slopes. Needless to say that more models are needed to confirm this possibility. Besides, SFs can potentially be used to study the turbulence eddies. Its flat profile at small scales may provide informations about the amplitude and size of the smallest turbulence cells. However, we showed that the results are affected by the observational resolution. \\subsection{Improved CF technique} The CF method has been proven an useful tool for the determination of the magnetic field in the ISM. We studied its validity using the obtained polarization maps from our models. \\cite{ostriker01} proposed that the CF method would only be applicable in restricted cases, in which the dispersion of the polarization angle is small ($< 25^{\\circ}$). It is consistent with the original approximations involving the derivation of the CF equation. We derived a generalized formula for the CF method, based on the same assumptions of the original work \\citep{chandra53}, but that accounts for larger dispersion models. Basically, we assume that the perturbations are Alfvenic and that we have an isotropic distribution of $\\delta V$. We, for the first time, successfully applied this equation to the super-Alfvenic and large inclination models. We also studied dependency of the CF method and the observational resolution. As also shown by \\cite{ostriker01} and \\cite{heitsch01}, the CF method overestimates the magnetic field for coarser resolutions. Therefore, we propose a general equation to fit the observational data considering maps with different resolution. The asymptotic value of the given procedure provides the ``infinite resolution\" measurement from the CF method and is consistent with the expected values from the simulations. As stated before, a possible limitation in the presented model is the absence of self-gravity effects. At the denser regions, the magnetic field configurations may possibly be different as the cloud collapses and drags the field lines. This process is responsible for the hour-glass structures observed in several gravitationally unstable clouds (e.g. Vall\\'ee \\& Fiege 2007). As a consequence, self-gravity increases the magnetic field locally and reduces the dispersion of the polarization angles within the dense clumps. To test the stability of the dense clumps in our simulations we may estimate the Jeans length ($\\lambda_J = c_s \\sqrt{\\pi/G\\rho}$). For the parameters chosen in Section 5.3, the denser structures, considering all models, are characterized by $l_{\\rm core} \\sim 0.1 - 0.5$pc, $n_{\\rm core} \\sim 10^{5-6}$cm$^{-3}$ and $T=10 - 100$K, resulting in $\\lambda_J \\sim 0.1 - 1$pc. Therefore, since $l_{\\rm core} \\sim \\lambda_J$, the denser structures may be unstable, at least for the given parameters. On the other hand, self-gravity plays a role at small regions and may not be statistically important for the previous results (except for the $P \\times I$ correlation) as we studied regions much larger than the very dense clumps. We plan to study the effects of self-gravity on the obtained results in a further work. \\subsection{Sub-Alfvenic versus super-Alfvenic turbulence} The ratio of thermal gas to magnetic pressures ($\\beta$) is typically used as the dominant parameter on the characterization of the degree of magnetization of a cloud. In this sense, systems with similar $\\beta$ values should present similar distribution of structures and dynamics. However, we showed that the sonic ($M_{\\rm S}$) and Alfvenic ($M_{\\rm A}$) Mach numbers, which quantify the ratio of the kinetic to the thermal and magnetic pressures, respectively, divide the models in different regimes. For the case of polarization vectors, our simulations showed that $M_{\\rm A}$ is decisive. For clouds with $M_{\\rm A} < 1$, the gas motions excited by turbulence are confined by the magnetic field and are not able to change its configuration. Actually, the perturbations in the magnetic field occur, but are small compared to the mean field ($\\delta B \\ll B_{\\rm m}$). In this case, the polarization vectors are uniform, as shown in Fig.\\ 4. For $M_{\\rm A} > 1$, the magnetic pressure is small compared to the kinetic energy of the turbulent gas and the mean magnetic field can be easily distorted. As a consequence, the polarization maps would show large dispersion of $\\phi$. Obviously, a large dispersion of $\\phi$ can also be related to a projection effect. If the mean magnetic field is projected along the line of sight ($\\theta \\sim 90\\degr$), only the small perturbations $\\delta B$ will be seen as the polarization vectors. However, considering a large number of clouds, there is a very low probability for all to present $\\theta \\sim 90\\degr$. In this case, if observations systematically show very large dispersions of $\\phi$, it means that the turbulence in the ISM may be typically super-Alfvenic. Otherwise, the ISM then presents sub(quasi)-Alfvenic turbulence. Observations of a given cloud could then be compared to our Fig.\\ 4 to determine under which regime the turbulence is operating. It is particularly interesting since the ratio of magnetic to turbulent energy in the ISM is still subject of controversy \\citep{padoan02, girart06}. \\subsection{Procedure for observational data analysis} The number of simulations presented in this work, as well as the numerical resolution, must be increased in future works. In any case, from the models we have, we intend to provide observers with a straightforward procedure to characterize the magnetic field and turbulence properties of molecular clouds. Firstly, from the polarization maps of a given region one should obtain the second order structure function of the polarization angle. From the SF, it is possible to characterize the turbulent cascade and the magnetic field. The extension of the flat profile at small scales give the turbulence cut-off scale. On the other extreme, the flat profile at large scales indicate the energy injection lengths. From the maximum slope of $SF \\propto l^{\\alpha}$, it is possible to determine the averaged Alfvenic Mach number and, as a consequence, the magnetic field intensity. Another method to obtain the magnetic field intensity is based on the CF technique. From the observed velocity dispersion it is possible to estimate the amplitude of the random component of the magnetic field from Eq.\\ (7). The total magnetic field is then obtained from Eq.\\ (9), using the dispersion of the polarization angle. To avoid the dependence on the observational resolution, it is suggested to evaluate the dispersion of the polarization angle for different resolutions (which may be simulated by averaging neighboring vectors of the polarization maps) and determine the asymptotic total magnetic field from Eq.\\ (10). Finally, subtracting the total field by the random component, it is possible to determine the mean magnetic field projected in the plane of sky. Also, combining the mean magnetic field obtained from both methods, it is possible to estimate the angle between the mean magnetic field and the LOS. \\subsection{Grain alignment} Although it was not considered in the present calculations, a correct treatment of grain alignment is needed for a full understanding of the polarization in molecular clouds. For instance, we showed that the polarization degree is anti-correlated with the column density with slope $\\gamma \\sim -0.5$, while observations sometimes give $\\gamma < -1.0$ \\citep{goncalves05}. This difference is related with the alignment efficiency at different regions of the cloud \\citep{cho05}. The theory of grain alignment has developed fastly during the past decade (see Lazarian [2007]). It is currently believed that radiative torques play a major role on the alignment process and it strongly depends on $A_V$ (see Lazarian \\& Hoang [2007]). With increasing extinction ($A_V > 2$), the radiative torques are less effective and only large grains are aligned. All in all, both observations \\citep{arce98, whittet08} and theory \\citep{hoang08}, suggest that there is a range of $A_V$ for which our assumptions are correct. It might happen that subsonic mechanical alignment of irregular grains, introduced in \\cite{lazho07b}, extends the range of $A_V$ over which grains are aligned when compared to the estimates based on radiative torques only. The observed band is also selective regarding the dust sizes and different bands reveal the polarization of different dust components. All these effects will be included in a future work, and a more realistic study of the polarization intensity distribution will be obtained. \\subsection{Polarization from molecular and atomic lines} In the present work we focused on calculating synthetic polarization maps of FIR emission from dust particles, which were assumed to be perfectly aligned with the magnetic field. Unfortunately, due to inefficient grain alignment at the dense cloud cores, the dust polarization degree may decrease and different methods have to be used. The polarization of molecular lines have been shown to be an additional tool for the study of the magnetic fields in the ISM \\citep{girart99,greaves02, girart04,cortes05}. Molecules are present in the dense and cold cores of the clouds and may be detected by thermal line emissions. Polarimetric maps of molecular emission can be used on the study of regions with $A_V > 10$, and can be directly associated with the Zeeman measurements. Based on the Goldreich-Kylafis effect \\citep{gold81,gold82}, the molecular sublevel populations will present imbalances due to the magnetic field, generating polarized rotational transitions. However, the survival of molecules depend on restrict conditions, as for $A_V > 10$ molecules may be frozen into dust particles. Another difficulty regarding this method is the fact that the GK effect generates polarization either parallel or perpendicular to the magnetic field lines, making the polarization maps. On the other hand, polarized scattering and absorption from atoms and ions provide information about the magnetic field in warm and rarefied regions, like the diffuse ISM and the intergalactic medium. Polarization arising from aligned atoms and ions is a new method \\citep{yan07a}. Unlike molecular lines that live in the excited state long enough to be imprinted by the magnetic field, the atomic exited states are short lived and tend to decay in timescales shorter than the Larmor precession of the atom. However, species with fine and/or hyperfine structure of ground or metastable states can be aligned. This fact opens new horizons for polarimetric studies of magnetic fields \\citep{yan06, yan07a, yan07b}. It is useful to comment here that the results shown in this work are also valid for the observed polarization maps of molecular and atomic emission lines. This because the assumptions made for the calculations disregard any special consideration about the emitting species, which could be atoms, molecules or dust particles. Also, it is worth mentioning that these techniques are either a substitute, for regions where no FIR dust emission is detectable, or complementary to the dust polarized emission but at different wavelengths (e.g.\\ optical and UV radiation). Depending on the $A_V$ range considered, polarization of atoms and molecules may complement the dust emission and absorption maps, which are usually much more detailed. \\subsection{Comparison with previous works} In this work we studied of polarization maps and its applicability on the determination of magnetic fields in molecular clouds based on numerical simulations. Here we compare the obtained results with the previous theoretical works. \\cite{ostriker01} performed numerical simulations, with $256^3$ resolution, considering plasma $\\beta$ values 0.01, 0.1 and 1.0. Their results showed homogeneous polarization maps for $\\beta = 0.01$ (strongly magnetized turbulence), and a complex distribution of polarization vector for $\\beta = 0.1$ and 1.0 (weakly magnetized turbulence). They also obtained an increase in the dispersion of polarization angles with the increase of the external magnetic field inclination regarding the line of sight. These results are in agreement with our models, except for the fact that two of our models with equal $\\beta$ presented completely different polarization maps. This because the $\\beta$ value does not reveal how the magnetic field lines respond to the turbulence. The Alfvenic and sonic Mach numbers reflect how strong is the turbulent pressure compared to the magnetic and thermal pressures, respectively. They obtained a higher polarization degree for $\\beta$ = 0.01, obviously because of the magnetic field intensity, but larger values of $P$ for larger column densities (i.e.\\ for larger $I$), in disagreement with observations. We believe that this was caused by their method for obtaining the polarization degree. They obtained the integrated Stokes parameters, weighted by local density, for all cells along the LOS. We, on the other hand, used a threshold on density to avoid the contribution of very rarefied regions (where the magnetic field structure is systematically more uniform). \\cite{padoan01} focused their work on the polarization of dust emission from dense cores. They implemented a more realistic calculation of the polarization degree based on the efficiency of the alignment for different values of $A_{\\rm V}$. In this case, they were able to obtain a decreasing polarization degree with the total intensity related to the grain properties, and not to the statistics of polarization vectors along the line of sight. We also believe that the numerical resolution may be playing a role on the polarization degree. More refined simulations systematically result in more complex structures for density and magnetic field lines. As a consequence, the alignment vectors along the line of sight present larger dispersion resulting in a lower polarization degree. We found no previous theoretical work presenting an extended statistical anaylsis considering all the PDF, Spectra and Structure Function of polarization angle, and therefore no comparison can be made. \\cite{ostriker01} and \\cite{padoan01} also tested the CF technique using their simulations, considering Eq.\\ (8). They obtained good agreement, with a correction factor of $\\sim 0.5$, between the calculated estimations and the expected values only for the models with $\\delta \\phi < 25\\degr$. \\cite{heitsch01} presented an extended analysis using a larger number of models, with different physical parameters and numerical resolutions (including 1 model with $512^3$ resolution). They also studied the effects of observational resolution on the obtained maps. They concluded that coarser resolutions result in more uniform polarization vectors. As a consequence, the CF method overestimates the magnetic field intensity. This is in full agreement with our results. They also tested the reliability of the CF technique in weakly magnetized clouds. They proposed the modified equation $B_{CF}B_{CF}^{mod}=4\\pi \\rho [\\delta v_{\\rm LOS}/\\delta(\\tan \\phi)][1+3\\delta(\\tan \\phi)^2]^{1/2}$, \\footnote{Here, $B_{CF}$ is the value obtained using the standard CF equation, and $B_{CF}^{mod}$ is the corrected value proposed by \\cite{heitsch01}.} which gave good results compared with the expected values for their models, with discrepancies of a factor $<2$. We tested their equation to our models 3 and 4 with $\\theta = 0$, representing a strong and weakly magnetized cloud, respectively. For Model 3, the obtained value is in agreement with that shown in Table 2. The ratio between the two measurements is $B_{CF}^{mod}/B_{\\rm CF}^0 = 0.9$. For Model 4, the proposed equation underestimates the magnetic field, and compared to with our method it gives $B_{CF}^{mod}/B_{\\rm CF}^0 \\sim 0.3$. Despite of the few simulations available for compariron, their method seems to systematically underestimate the magnetic field intensity. More tests are needed to determine which method may give the best results. Furthermore, even though not addressed by \\cite{heitsch01}, we studied the dependence of the polarization angles and the CF technique with the inclination of the magnetic field regarding the LOS. We showed that there is a degeneracy between the results of weakly magnetized clouds and strongly magnetized clouds with high $\\theta$. The modified CF formula presented in this work gave good results for all cases. As discussed before, even though not taking into account the self-gravity in our simulations it mostly induces changes at small scales, as noted by \\cite{heitsch01}. As a result, they showed the CF technique to be insensitive to self-gravity." }, "0801/0801.2560_arXiv.txt": { "abstract": "Terrestrial planets form in a series of dynamical steps from the solid component of circumstellar disks. First, km-sized planetesimals form likely via a combination of sticky collisions, turbulent concentration of solids, and gravitational collapse from micron-sized dust grains in the thin disk midplane. Second, planetesimals coalesce to form Moon- to Mars-sized protoplanets, also called ``planetary embryos''. Finally, full-sized terrestrial planets accrete from protoplanets and planetesimals. This final stage of accretion lasts about 10-100 Myr and is strongly affected by gravitational perturbations from any gas giant planets, which are constrained to form more quickly, during the 1-10 Myr lifetime of the gaseous component of the disk. It is during this final stage that the bulk compositions and volatile (e.g., water) contents of terrestrial planets are set, depending on their feeding zones and the amount of radial mixing that occurs. The main factors that influence terrestrial planet formation are the mass and surface density profile of the disk, and the perturbations from giant planets and binary companions if they exist. Simple accretion models predicts that low-mass stars should form small, dry planets in their habitable zones. The migration of a giant planet through a disk of rocky bodies does not completely impede terrestrial planet growth. Rather, \"hot Jupiter\" systems are likely to also contain exterior, very water-rich Earth-like planets, and also \"hot Earths\", very close-in rocky planets. Roughly one third of the known systems of extra-solar (giant) planets could allow a terrestrial planet to form in the habitable zone. ", "introduction": "Recent research has developed a model for the growth of the terrestrial planets in the Solar System via collisional accumulation of smaller bodies (e.g., Wetherill 1990). Several distinct dynamical stages have been identified in this process, from the accumulation of micron-sized dust grains to the late stage of giant impacts. In this article, I first review the current state of knowledge of the stages of terrestrial planet formation ($\\S$ 2). Second, I explore the effects of external parameters on the accretion process, including new work showing that simple accretion models predict that low-mass stars should preferentially harbor low-mass terrestrial planets in their habitable zones ($\\S$ 3). These include the orbit of a giant planet and the surface density profile of the protoplanetary disk. Third, I examine the effects of the migration of a Jupiter-mass giant planet on the accretion of terrestrial planets ($\\S$ 4), and also apply accretion models to the known set of extra-solar planets to show that roughly one third of the known systems could have formed an Earth-like planet in the habitable zone ($\\S$ 5). Finally, conclusions and avenues for future study are presented ($\\S$ 6). ", "conclusions": "In this proceedings, I have summarized the current state of knowledge about terrestrial planet formation in extra-solar planetary systems. In $\\S$ 2, I reviewed the stages of terrestrial planet formation, from micron-sized grains to planetesimals, protoplanets, and full-sized planets. In $\\S$ 3 I described the effect of external parameters such as the disk and giant planet properties on the terrestrial accretion process as well as radial mixing and water delivery. In $\\S$ 4, I presented recent results showing that terrestrial planets can form in systems with close-in giant planets, assuming those to have migrated to their final locations. In fact, close-in giant planets should be accompanied by both very close-in \"hot Earths\" and exterior ocean planets (Raymond \\etal 2006b). In $\\S$ 5, I used results from previous work to derive limits designed to predict, based on the observed orbits of giant planets, where to search for other Earths among the known extra-solar systems. Approximately one third of the known systems of giant planets are good candidates for harboring an Earth-like planet (Mandell \\etal 2007). Many issues remain to be resolved in each topic I described. The details of planetesimal formation are still poorly known, although recent results suggest that several process including turbulence and migration of meter-sized bodies acting in tandem might be the solution (Johansen \\etal 2007). The details and consequences of giant impacts are not well understood, in terms of the fate of collisional debris and compositional changes induced by the impacts (Genda \\& Abe 2005; Asphaug \\etal 2006; Canup \\& Pierazzo 2006). In addition, some of the effects of external parameters are perhaps less well understood than we would like to think. For example the giant planet eccentricity - terrestrial planet water content correlation (Chambers \\& Cassen 2002; Raymond \\etal 2004) inherently assumes that eccentric giant planets acquire their eccentricities early, before terrestrial accretion. If this assumption is wrong, then one can imagine a scenario in which terrestrial planets tend to form with relatively circular giant planets; this is a beneficial situation in terms of water delivery. However, a late, impulsive eccentricity increase can destabilize the orbits of terrestrial planets and remove them from the system entirely (Veras \\& Armitage 2006). Indeed, there are many known extra-solar planets that would be favorable for terrestrial planet formation if their orbits were more circular (see Fig.~\\ref{fig:xsp}). Thus, if giant planet eccentricity is acquired late, many systems may undergo ``planetary system suicide'', forming an Earth-like planet and subsequently destroying it. A more complete, holistic view of planet formation and evolution is needed to distentangle these effects." }, "0801/0801.0423_arXiv.txt": { "abstract": "{} {A recent catalogue by Flesch \\& Hardcastle presents two major anomalies in the spatial distribution of QSO candidates: $i/$ an apparent excess of such objects near bright galaxies, and $ii/$ an excess of very bright QSO candidates compared to random background expectations in several regions of the sky. Because anyone of these anomalies would be relevant in a cosmological context, we carried out an extensive analysis of the probabilities quoted in that catalogue.} {We determine the nature and redshift of a subsample of 30 sources in that catalogue by analysing their optical spectra (another 11 candidates were identified from existing public databases). These have allowed us to statistically check the reliability of the probabilities QSO status quoted by Flesch \\& Hardcastle for their candidates.} {Only 12 of the 41 candidates turned out QSOs (7 of which have been identified here for the first time).} {The probabilities of the QSOs' being the candidates given by Flesch \\& Hardcastle are overestimated for $m_B\\le 17$ and for objects projected near ($\\le 1$ arcmin) bright galaxies. This is the cause of the anomalies mentioned above.} ", "introduction": "At present, statistics on QSO distributions can be carried out using the recent public compilations of QSOs identified in the SDSS (Adelman-McCarthy et al.\\ 2007) and 2dF (Croom et al.\\ 2004) surveys. Furthermore, Flesch \\& Hardcastle (2004; hereafter FH04) present an all-sky catalogue with 86\\,009 optical counterparts of radio/X-ray sources as QSO candidates (with a probability greater than 40\\%, according to FH04) that were not identified previously as such. This could be a useful catalogue to look for QSOs in regions that were not covered by SDSS or 2dF, or for deeper magnitudes than their completeness limit. Nevertheless, since the FH04 catalogue does not give the identification of each source but just a statistical estimate of the probability that it be a QSO (or a star, a galaxy, or a wrongly identified optical counterpart of the X-ray/radio source), it must be used with care. Flesch\\footnote{See his web-page http://quasars.org/ .} claims that his catalogue `reveals a few galaxy-centered fields which are populated by QSOs/candidates more thickly than expected by the background density of such QSOs,' which would argue in favour of the hypothesis of QSOs with anomalous redshift and at the same distance as the galaxies with which they are associated (Arp 1998, Burbidge 2001). Because this is against one of the most basic assumptions in standard modern cosmology, we consider it interesting to check it on the basis of rigorous statistical computation. This is the main objective of this paper. In \\S \\ref{.excesses}, we analyse in which cases (on the basis of the probabilities quoted by FH04) there are anomalies in the distribution of QSOs. By conducting optical spectroscopy (\\S \\ref{.select}), we directly check the nature of a subsample of FH04 sources. On the basis of these observations, in \\S \\ref{.probab} we check the reliability of the probabilities quoted by FH04. In \\S \\ref{.conclusions} we reanalyse the distribution of QSOs and discuss the implications of our results. ", "conclusions": "\\label{.conclusions} In FH04 there are 88 different objects (106 cases with repetitions) with $m_B\\le 18$ within a radius of 3 arcminutes of RC3 galaxies. Of these 88 objects, there are spectroscopic observations of the optical counterparts in 26 cases (36 cases including repetitions), and 10 of these (the same number with repetitions) are in fact QSOs (while the 16 other objects (26 with repetitions) are stars, galaxies, or HII regions). The average expected number of background cases around all RC3 cases should be $\\approx 70$ (\\S \\ref{.explorc3}) and, in our subsample, should be $\\approx \\frac{36}{106}\\times 70=24$ cases, which is higher than the value derived from the observations. This stems from the incompleteness of the FH04 catalogue. Within a radius of 10 arcminutes of RC3-galaxies, there are 14 objects (no repetitions) with $m_B\\le 15.5$, of which five (36\\% of the sample) were observed and only one (\\# 5 with $m_B=15.4$) turned out a QSO. The average expected number of cases around all RC3 galaxies is 7.6 (\\S \\ref{.explorc3}), and in our subsample should be $\\frac{5}{14}\\times 7.6=2.7$ cases. This result agrees with expectations from a background distribution of QSOs. Limiting the study to RC3 galaxies by $m_{B,gal}\\le 14.5$, $m_{B,gal}\\le 13.5$, and $m_{B,gal}\\le 12.5$, the number of galaxies is reduced to 7,675/2,566/883. For the subsample with $m_b\\le 18$, $d\\le 3'$ we obtained 5/2/2 QSOs to be compared with the expected 7.8/2.7/0.92. For the subsample with $m_b\\le 15.5$, $d\\le 10'$, we obtained 1/0/0 to be compared with 0.50/0.17/0.06. There is no statistical inconsistency with the claim that all QSOs are background QSOs rather than being associated with the galaxies. As for the distribution of the QSOs with magnitude or distance from RC3 galaxies, as pointed out in \\S \\ref{.probab}, there is no longer any inconsistency with background predictions: almost all the confirmed QSOs in our subsample are faint (all of them with magnitudes $m_B\\ge 17.0$, except \\#5) and distant enough from the centres of RC3 galaxies to be compatible with background expectations. See Fig. \\ref{Fig:QSOdist2} for the distance distribution and note how the observed QSOs are fewer than the background expectations for all distances. Some of the most troublesome candidates, such as those shown in Fig. \\ref{Fig:examples}, were in the end confirmed to be objects different from QSOs. The only QSO with $m_B\\le 15.5$ (\\# 5) within 10$'$ from an RC3 galaxy is associated with NGC 1136, which was previously reported as a galaxy with peculiar associations (Arp 1981). Objects claimed by Arp to be anomalous are much fewer than 23,011 RC3 galaxies. They are $\\sim 10^2$, so the probability of finding a QSO with $m_B\\le 15.5$ within 10$'$ from some of them is $P\\sim 0.06$. It is not few enough to claim it separately from other facts as a possible anomalous case, but a fact to be added in the discussion of the case of NGC 1136. For the object \\# 6, now confirmed to be a QSO with $m_B=16.5$ at only 30$\"$ from PHL 1459 (a galaxy with $m_B=17.2$ and X-ray emission), we are not within the statistics of RC3 galaxies. There are $\\approx 4.3\\times 10^5$ galaxies in the whole sky up to the magnitude $m_B=17.2$ (Metcalfe et al.\\ 1991), so, using Eq.\\ (\\ref{densQSO}), the expected number of cases like this within a 30-arcsecond distance of these galaxies is 2.2. The discovery of only one case is not enough to claim any statistical anomaly. If we take into account the peculiarity that PHL 1459 is an X-ray source and has $\\approx 120$ counts/hour in ROSAT (Voges et al.\\ 1999, 2000), the probability is somewhat lower. There are $\\sim 0.4$ galaxies/deg$^2$ up to this X-ray flux (Voges et al. 1999; Zickgraf et al. 2003), i.e. around 16,000 galaxies like this in the whole sky. The probability of finding a QSO with $m_B=16.5$ within 30$\"$ of one of them is 0.09; not low enough to discard a background projection coincidence. \\begin{figure} \\vspace{1cm} \\resizebox*{7cm}{7cm}{\\includegraphics{QSOdist2.eps}}\\par \\caption{Cumulative counts of QSOs of this paper up to magnitude $m_B=18$ [34\\% (36/106) of the QSO/RC3--galaxy pairs] as a function of the maximum distance to the galaxy. QSOs associated with two galaxies count twice with each corresponding distance.} \\label{Fig:QSOdist2} \\end{figure} Summing up, within the subsample of 41 objects from FH04, we have not found any statistical anomaly with respect to the expectation of QSOs as background sources. Since we observed a representative subsample of the most controversial cases that could in principle present some excesses of QSOs, we think that there should be in principle no reason to expect statistical anomalies in the FH04 catalogue, which is anyway a useful database for looking for QSOs in regions not covered by other surveys, such as SDSS or 2dF, or for deeper magnitudes. Nevertheless, there are still anomalies in the QSO distributions presented with some other catalogues (e.g.\\ Arp 1998; Burbidge 2001) that require further attention. Further rigorous statistical analysis like those presented here and in L\\'opez-Corredoira \\& Guti\\'errez (2006$a,b$) are necessary. \\ {\\bf Acknowledgments:} Thanks are given to Rub\\'en J. D\\'\\i az (Gemini Observatory) for providing telescope time at 2.15 m CASLEO, to Eric Flesch for his comments and criticism on a draft of this paper, to the referee Mira Veron for her helpful comments for improving it, to T. J. Mahoney (IAC) and Joli Adams (A\\&A) for proofreading the paper. Based on observations made with the telescopes: 2.15 m CASLEO in the observatory of ``El Leoncito'', San Juan, Argentina; 1.93 m OHP (these observations were funded by the Optical Infrared Coordination network, OPTICON, a major international collaboration supported by the Research Infrastructures Programme of the European Commission Sixth Framework Programme), Haute-Provence, France; and 2 m IUCAA (Pune, India). This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France. The Digitized Sky Surveys were produced at the Space Telescope Science Institute under US 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. Funding for the creation and distribution of the SDSS Archive has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Aeronautics and Space Administration, the National Science Foundation, the U.S. Department of Energy, the Japanese Monbukagakusho, and the Max Planck Society. The SDSS Web site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium (ARC) for the Participating Institutions. The Participating Institutions are The University of Chicago, Fermilab, the Institute for Advanced Study, the Japan Participation Group, The Johns Hopkins University, the Korean Scientist Group, Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University of Washington. The first author (MLC) was supported by the {\\it Ram\\'on y Cajal} Programme of the Spanish Science Ministry." }, "0801/0801.2432_arXiv.txt": { "abstract": "We have performed a joint analysis of prompt emission from four bright short gamma-ray bursts (GRBs) with the Suzaku-WAM and the Konus-Wind experiments. This joint analysis allows us to investigate the spectral properties of short-duration bursts over a wider energy band with a higher accuracy. We find that these bursts have a high E$_{\\rm peak}$, around 1 MeV and have a harder power-law component than that of long GRBs. However, we can not determine whether these spectra follow the cut-off power-law model or the Band model. We also investigated the spectral lag, hardness ratio, inferred isotropic radiation energy and existence of a soft emission hump, in order to classify them into short or long GRBs using several criteria, in addition to the burst duration. We find that all criteria, except for the existence of the soft hump, support the fact that our four GRB samples are correctly classified as belonging to the short class. In addition, our broad-band analysis revealed that there is no evidence of GRBs with a very large hardness ratio, as seen in the BATSE short GRB sample, and that the spectral lag of our four short GRBs is consistent with zero, even in the MeV energy band, unlike long GRBs. Although our short GRB samples are still limited, these results suggest that the spectral hardness of short GRBs might not differ significantly from that of long GRBs, and also that the spectral lag at high energies could be a strong criterion for burst classification. ", "introduction": "The bimodal distribution of the duration of gamma-ray bursts (GRBs) indicates that there are two distinct classes of events. The long-duration bursts have typical durations of around 20 s, while the short-duration bursts (about 1/4 of the total) have durations of around 0.3 s (\\cite{Mazets 1981}; \\cite{Norris 1984}; \\cite{Dezalay 1992}; \\cite{Hurley 1992}; \\cite{Kouveliotou 1993}; \\cite{Norris 2000}). This suggests that the short and long GRBs may have different progenitors. Core collapse of massive stars and mergers of compact binaries are considered likely models for the long and short GRBs, respectively (\\cite{Katz_Canel}; \\cite{Ruffert 1999}). Recently, thanks to the rapid position information provided by HETE-2 (\\cite{Rikker 2003}) and Swift (\\cite{Gehrels 2004}), afterglow observations have progressed dramatically and X-ray and optical afterglow emissions were discovered from some short GRBs (\\cite{Hjorth 2005}; \\cite{Fox 2005}), as well as long ones. Some short GRB afterglows have been found to be associated with galaxies not undergoing star formation, while some long GRBs have been found to be associated with energetic supernovae. These results support the hypothesis that long and short GRBs indeed have different progenitors. From the point of view of prompt gamma-ray emission, although the spectral characteristics of long GRBs have been well studied (\\cite{Frontera 2000}: BeppoSAX; \\cite{Kaneko 2006}: BATSE; \\cite{Sakamoto 2005b}: HETE-2), our understanding of that of short GRBs is still incomplete, in part due to their very short durations. The spectral characteristics of BATSE short GRBs are often characterized by the hardness ratio, that is, the ratio of 100--300 keV to 25--100 keV counts or fluence (\\cite{Kouveliotou 1993}; \\cite{Cline 1999}). These studies suggest that, although there is considerable overlap, short GRBs tend to be harder than long GRBs. Paciesas et al. (2001) and Ghirlanda et al. (2004) compared the spectral parameters of bright BATSE short GRBs with those of long GRBs by spectral fitting. They pointed out that the spectra of short GRBs are well described by a cut-off power-law model and that the Band model (\\cite{Band 1993}) did not improve the fit. They also confirmed that the spectra of short GRBs are harder than those of long GRBs. This was found to be a consequence of a flat low energy photon index, rather than a difference in the peak energies. There is another complicating issue in the classification of short and long GRBs. Since their duration distributions overlap, other distinguishing criteria have been proposed, such as differences in the host galaxy (\\cite{Hjorth 2003}; \\cite{Hjorth 2005}), spectral hardness (\\cite{Cline 1999}), spectral lag (\\cite{Norris 2002}; \\cite{Norris 2006}), isotropic radiation energy (\\cite{Amati 2002}; \\cite{Amati 2006}), and the existence of a soft hump (\\cite{Norris 2006}), or a combination of these and other characteristics (\\cite{Donaghy 2006}). Indeed, some GRBs cannot be classified as short or long using the burst duration alone. For example, GRB 040429 was classified as short using the burst duration (2.39 s), but other properties such as spectral lag and hardness resembled those of long GRBs (\\cite{Wiersema 2004}; \\cite{Fox_Moon 2004}; \\cite{Donaghy 2006}). GRB 051227 (\\cite{Hullinger 2005}; \\cite{Barthelmy 2005a}; \\cite{Sakamoto 2005}) had a long enough duration of 8.0 s to be classified as long. However, this burst had no significant spectral lag in the initial spike, which is seen in many long GRBs (\\cite{Norris 1996}), and it also had a long soft hump in the light curve. These two properties are similar to those of short GRBs. GRB 060614 (\\cite{Gehrels 2006}) exhibited a long duration (102 s). However, it did not have any significant spectral lag in the initial spike, and also there were no signs of any associated supernova, despite its distance of z = 0.125, which should have been close enough to detect it. These examples show why the burst duration should no longer be considered a sole indicator for distinguishing between the short and long classes of bursts. Indeed Donaghy et al. (2006) have suggested the terms ``short population bursts'' and ``long population bursts'' to distinguish the two classes. In this paper, we will use simply ``short'' and ``long'' to describe these classes. Here, we report on the joint spectral analysis of four bright short GRBs simultaneously observed by the Suzaku-WAM (Suzaku Wide-band All-sky Monitor) and Konus-Wind. The WAM and Konus have wide energy ranges, from 50--5000 keV and 10--10000 keV, respectively, and the WAM has the largest effective area from 300 keV to 5000 keV of any current experiment with spectral capabilities. Thus, we can investigate the spectral characteristics of short GRBs up to MeV energies. We also classify the four GRBs as short or long using our data with high statistics up to the MeV energy region using a number of criteria: spectral lag, hardness ratio, isotropic radiation energy and existence of a soft hump. ", "conclusions": "We performed a joint analysis of four bright, hard-spectrum, short-duration GRBs, localized by the IPN, with the Suzaku-WAM and Konus-Wind, in order to investigate the spectral properties of short GRBs into the MeV energy band. From the spectral analysis, we found that these bursts have a high E$_{\\rm peak}$ around 1 MeV; the spectral parameters can be constrained tightly by joint fitting. However, we could not determine whether these spectra follow the CPL model or the Band model, because there is no significant improvement between these models. The power-law photon index is slightly flatter than that of long BATSE GRBs. In particular, GRB 051127 has a very flat spectrum with a photon index, $\\alpha$, that is larger than the synchrotron limit of $-$2/3. We also examined several other spectral and temporal properties to confirm that the bursts belong to the short class. Most properties, such as the spectral lag, the spectral hardness, and the isotropic radiation energy, indicated that the four bursts were indeed in this class. In addition, a broad-band spectral analysis with the WAM and Konus revealed some interesting properties of short GRBs: i) There is no evidence for very hard GRBs that have the large hardness ratios seen in BATSE short GRBs. ii) The spectral lag of our short GRBs does not show any clear evolution, even in the MeV energy region, unlike that of long hard GRBs. Although our short GRB sample is still limited, these results might suggest one possibility, that the hardness ratios of the short GRBs are not so different from those of long GRBs, and also that the spectral lag at high energies can be a strong criterion for classifying short and long GRBs. \\bigskip M. O. is supported by the Research Fellowships of the Japan Society for the Promotion of Science for Young Scientists (2006). K. H. is grateful for IPN support under the NASA LTSA program, grant FDNAG5-11451, the INTEGRAL Guest Investigator program, NAG5-12706 and NNG06GE69G, and the Suzaku Guest Investigator program NNX06AI36G. The Konus-Wind experiment is supported by a Russian Space Agency contract and RFBR grant 06-02-16070. We are grateful to Richard Starr for providing the Mars Odyssey data, and to Giselher Lichti, Arne Rau, and Andreas von Kienlin for providing the INTEGRAL SPI-ACS data. We would like to thank Jay Norris for useful advice concerning the spectral lag analysis." }, "0801/0801.3744_arXiv.txt": { "abstract": "Observations show that small-amplitude prominence oscillations are usually damped after a few periods. This phenomenon has been theoretically investigated in terms of non-ideal magnetoacoustic waves, non-adiabatic effects being the best candidates to explain the damping in the case of slow modes. We study the attenuation of non-adiabatic magnetoacoustic waves in a slab prominence embedded in the coronal medium. We assume an equilibrium configuration with a transverse magnetic field to the slab axis and investigate wave damping by thermal conduction and radiative losses. The magnetohydrodynamic equations are considered in their linearised form and terms representing thermal conduction, radiation and heating are included in the energy equation. The differential equations that govern linear slow and fast modes are numerically solved to obtain the complex oscillatory frequency and the corresponding eigenfunctions. We find that coronal thermal conduction and radiative losses from the prominence plasma reveal as the most relevant damping mechanisms. Both mechanisms govern together the attenuation of hybrid modes, whereas prominence radiation is responsible for the damping of internal modes and coronal conduction essentially dominates the attenuation of external modes. In addition, the energy transfer between the prominence and the corona caused by thermal conduction has a noticeable effect on the wave stability, radiative losses from the prominence plasma being of paramount importance for the thermal stability of fast modes. We conclude that slow modes are efficiently damped, with damping times compatible with observations. On the contrary, fast modes are less attenuated by non-adiabatic effects and their damping times are several orders of magnitude larger than those observed. The presence of the corona causes a decrease of the damping times with respect to those of an isolated prominence slab, but its effect is still insufficient to obtain damping times of the order of the period in the case of fast modes. ", "introduction": "\\label{sec:intro} Solar prominences are large-scale coronal magnetic structures whose material, cooler and denser than the typical coronal medium, is in plasma state. Prominences are supported against gravity by the coronal magnetic field, which also maintains the prominence material thermally isolated from the corona. Small-amplitude oscillations in solar prominences were detected almost 40 years ago \\citep{harvey}. These oscillatory motions seem to be of local nature and their velocity amplitude is typically less than 2--3~km~s$^{-1}$. Observations have also allowed to measure a wide range of periods between 30~s \\citep{balthasar} and 12~h \\citep{foullon}. More recently, some high-resolution observations of prominence oscillations by the Hinode/SOT instrument have been reported \\citep{okamoto,berger,ofman}. From the theoretical point of view, the oscillations have been interpreted by means of the magnetoacoustic eigenmodes supported by the prominence body. A recent example is the work by \\citet{hinode} in which the observations of \\citet{okamoto} are interpreted as fast kink waves. The reader is referred to \\citet{oliverballester02,ballester,banerjee} for extensive reviews of both observational and theoretical studies. Evidence of the attenuation of small-amplitude prominence oscillations has been reported in some works \\citep{molowny99,terradasrad,lin2004}. A typical feature of these observations is that the oscillatory motions disappear after a few periods, hence they are quickly damped by one or several mechanisms. The theoretical investigation of this phenomenon in terms of magnetohydrodynamic (MHD) waves has been broached by some authors by removing the ideal assumption and by including dissipative terms in the basic equations. Non-adiabatic effects appear to be very efficient damping mechanisms and have been investigated with the help of simple prominence models \\citep{ballai,carbonell,spatial,terradas}. Nevertheless, other damping mechanisms have been also proposed, like wave leakage \\citep{schutgensA,schutgensB,schutgensToth}, dissipation by ion-neutral collisions \\citep{forteza} and resonant absorption \\citep{arregui}. In a previous work \\citep[][hereafter Paper~I]{soler}, we have studied for the first time the wave attenuation by non-adiabatic effects of a prominence slab embedded in the corona. In that work the magnetic field is parallel to the slab axis and it is found that the corona has no influence on the internal slow modes, but it is of paramount importance to explain the damping of fast modes, which are more attenuated than in simple models that do not consider the coronal medium. Following the path initiated in Paper~I, here we investigate the wave damping due to non-adiabatic mechanisms (radiative losses and thermal conduction) in an equilibrium made of a prominence slab embedded in a coronal medium, but now we consider a magnetic field transverse to the slab axis. This configuration and that studied in Paper~I correspond to limit cases, since measurements with Zeeman and Hanle effects indicate that the magnetic field lines are skewed to the long axis of prominences. On average, the prominence axis and the magnetic field form an angle of about 20~deg. Thus, the skewed case is relegated to a future investigation. The equilibrium configuration assumed here was analysed in detail by \\citet{JR92} and \\citet{oliver} in the case of ideal, adiabatic perturbations. The main difference between both works is in the treatment of gravity. \\citet{JR92} neglected the effect of gravity and so straight field lines were considered. On the other hand, \\citet{oliver} took gravity into account and assumed curved field lines according to the \\citet{kippen} model modified to include the surrounding coronal plasma \\citep{poland}. Despite this difference, both studies agree in establishing a distinction between different normal modes depending on the dominant medium supporting the oscillation. Hence, internal modes are essentially supported by the prominence slab whereas external modes arise from the presence of the corona. In addition, hybrid (or string) modes appear due to the combined effect of both media. The investigation of the thermal attenuation of oscillations supported by such equilibrium is unsettled to date and, indeed, this is the main motivation for the present study. However, two works \\citep{terradasrad01,terradas} studied the wave damping in an isolated prominence slab. \\citet{terradasrad01} considered radiative losses given by the Newtonian law of cooling as damping mechanism and studied the attenuation in the \\citet{kippen} and \\citet{menzel} prominence models. Subsequently, \\citet{terradas} considered a more complete energy equation including optically thin radiation, plasma heating and parallel thermal conduction, and assumed straight field lines since gravity was neglected. The main conclusion of both works is that non-adiabatic mechanisms are only efficient in damping slow modes whereas fast modes remain almost undamped. Nevertheless, in the light of the results of Paper~I, the presence of the coronal medium can have an important repercussion on the wave damping. The investigation of this effect is the main aim of the present work. Therefore, we extend here the work of \\citet{terradas} by considering the presence of the corona and neglect the effect of gravity as in \\citet{JR92} for simplicity. This paper is organised as follows. Section~\\ref{sec:equilibrium} contains a description of the equilibrium configuration and the basic equations which govern non-adiabatic magnetoacoustic waves. Then, the results of this work are extensively discussed in Sect.~\\ref{sec:results}. Finally, our conclusions are given in Sect.~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} In this paper we have studied the wave attenuation in a system representing a quiescent solar prominence embedded in the coronal medium. The prominence has been modelled as a homogeneous plasma slab surrounded by a homogeneous medium with coronal conditions. Magnetic field lines have been assumed transverse to the prominence slab axis and the whole system has been bounded in the field direction by two photospheric rigid walls, in order to establish a realistic length for the field lines. The attenuation of the normal modes of such equilibrium has been investigated by considering parallel thermal conduction, radiative losses and plasma heating as non-adiabatic mechanisms, and focusing our study on the fundamental oscillatory modes. The main conclusions of this work are summarised next. \\begin{enumerate} \\item Slow modes are strongly attenuated by non-adiabatic mechanisms, their damping times being of the order of the corresponding periods. Fast modes are less affected and present greater damping times. \\item The most relevant damping mechanisms are prominence radiation and coronal thermal conduction. The first one dominates the damping of internal modes, while the second one is responsible for the attenuation of external modes. The combined effect of both mechanisms governs the damping of hybrid modes. Neither prominence conduction nor coronal radiation become of importance for realistic values of the length of magnetic field lines and the prominence width. \\item The attenuation of slow modes is not affected by the value of the free component of the wavenumber, $k_z$. On the contrary, the behaviour of fast modes is strongly dependent on $k_z$. \\item Thermal conduction allows energy transfer between the prominence slab and the coronal medium. Prominence radiation has an essential role in dissipating the extra heat injected from the corona and stabilises the oscillations. Thermal instabilities appear if the radiative losses from the prominence plasma are omitted or significantly reduced (e.g. caused by an increase of the optical thickness) since the plasma cannot dissipate the extra injected heat in an efficient way. \\item The damping time of fast modes is strongly sensitive to the equilibrium physical parameters while slow waves are less affected by the variation of the equilibrium conditions. \\item The presence of the corona produces a decrement of the damping time of internal modes with respect to the solutions supported by an isolated prominence slab. Nevertheless, this effect is not enough to obtain damping times of the order of the period in the case of fast modes. \\end{enumerate} Considering the equilibrium parameters of Paper~I, the efficiency of non-adiabatic mechanisms on the damping of fast modes is smaller in the present case. This fact suggests that the orientation of magnetic field lines with respect to the slab axis has a relevant influence on the attenuation of fast modes, the configuration of Paper~I and the present one being limit cases. Moreover, fast modes are strongly sensitive to the equilibrium physical conditions, and it is possible to obtain small values of the damping time by considering extreme equilibrium parameters, such as very weak magnetic fields and very large prominence densities. In this way, fast modes show a wide range of theoretical damping times. On the other hand slow modes are always efficiently attenuated, with damping times of the order of their periods. This result suggests that the attenuation of prominence fast waves may be caused by other damping mechanisms not considered here. Some candidates could be resonant absorption \\citep{arregui} and ion-neutral collisions \\citep{forteza}. Among these mechanisms, resonant absortion may be a very efficient damping mechanism if non-uniform equilibria are considered, e.g. models with a transition region between the prominence and the corona. Other effects, as wave leakage, might only play a minor role in the damping of disturbances. Finally, future studies should take into account the prominence fine structure on the basis that small-amplitude oscillations are of local nature. Therefore, the investigation of the damping of fibril oscillations should be the next step. R.~O. and J.~L.~B. want to acknowledge the International Space Science Institute teams ``Coronal waves and Oscillations'' and ``Spectroscopy and Imaging of quiescent and eruptive solar prominences from space'' for useful discussions. The authors acknowledge the financial support received from the Spanish MCyT and the Conselleria d'Economia, Hisenda i Innovaci\\'o of the CAIB under Grants No. AYA2006-07637 and PCTIB-2005GC3-03, respectively. Finally, R.~S. thanks the Conselleria d'Economia, Hisenda i Innovaci\\'o for a fellowship." }, "0801/0801.4955_arXiv.txt": { "abstract": "We present the results of a systematic study of mid-IR spectra of Galactic regions, Magellanic \\hii\\ regions, and galaxies of various types (dwarf, spiral, starburst), observed by the satellites \\iso\\ and \\spitz. We study the relative variations of the 6.2, 7.7, 8.6 and $11.3\\mic$ features inside spatially resolved objects (such as \\M{82}, \\M{51}, \\xxxdor, \\M{17} and the \\orb), as well as among 90 integrated spectra of 50 objects. Our main results are that the 6.2, 7.7 and $8.6\\mic$ bands are essentially tied together, while the ratios between these bands and the $11.3\\mic$ band varies by one order of magnitude. This implies that the properties of the PAHs are remarkably universal throughout our sample, and that the relative variations of the band ratios are mainly controled by the fraction of ionized PAHs. In particular, we show that we can rule out both the modification of the PAH size distribution, and the mid-infrared extinction, as an explanation of these variations. Using a few well-studied Galactic regions (including the spectral image of the \\orb), we give an empirical relation between the $\\ipah{6.2}/\\ipah{11.3}$ ratio and the ionization/recombination ratio $G_0/n_e\\times\\sqrt{T_\\sms{gas}}$, therefore providing a useful quantitative diagnostic tool of the physical conditions in the regions where the PAH emission originates. Finally, we discuss the physical interpretation of the $\\ipah{6.2}/\\ipah{11.3}$ ratio, on galactic size scales. ", "introduction": "\\label{sec:intro} The reprocessing of stellar light by dust in the infrared (IR) is widely used to probe embedded star formation. However, in the absence of other constraints, the physical properties which can usually be derived from an almost-featureless grain continuum emission is limited to global quantities, such as the dust mass and its average temperature. The ubiquity of numerous mid-IR aromatic features, in a wide variety of astrophysical objects and environments, potentially provides more articulate diagnostics of the physical conditions. Indeed, these features dominate the mid-IR spectra of evolved stars \\citep[{e.g.}][]{blommaert05,kraemer06}, the cool ISM \\citep[{e.g.}][]{abergel05,flagey06,povich07}, as well as whole galaxies \\citep[{e.g.}][]{verma05,smith07}, that have been extensively observed by the \\isoST\\ (\\iso), and are currently investigated with a higher sensitivity by the \\spitzST. In our Galaxy, one third of the stellar light is reprocessed by dust, while this fraction can go up to $99\\%$ and higher, in starburst galaxies. At solar metallicity, roughly $15\\%$ of the cooling is radiated through the most powerful mid-IR bands, centered at 3.3, 6.2, 7.7, 11.3 and 12.7$\\mic$. Historically, these emission features were attributed to very small grains ($\\simeq10\\,$\\AA), transiently heated by single photon absorption, in order to account for the independence of the color temperature with the distance from the illuminating star, in several reflection nebulae \\citep{sellgren84}. In parallel, the central wavelengths of these bands were recognized to coincide with the vibrational modes of aromatic material \\citep{duley81}. These features are now commonly attributed to the molecular modes of Polycyclic Aromatic Hydrocarbons \\citep[hereafter PAHs;][]{leger84,allamandola85,allamandola89}, which are planar molecules made of $\\simeq10$ to 1000 carbon atoms, excited primarily by ultraviolet (UV) photons. With silicate and carbon grains, PAHs are a main component of dust models \\citep{desert90,dwek97,draine01,zubko04}. Their absorption efficiency has been modeled using astrophysical observations, laboratory measurements and quantum theory \\citep[in particular][]{desert90,joblin92,verstraete01,li01,mattioda05,mattioda05b,draine07,malloci07}. In addition to being major radiative coolants of the interstellar medium (ISM), PAHs are responsible for most of the photoelectric heating of the gas in photodissociation regions (hereafter PDRs) and the neutral interstellar medium, due to their high cumulative surface area \\citep[{e.g.}][]{bakes94,hollenbach97}. For the same reason, they probably play an important role in grain surface chemistry \\citep[{e.g.}][]{tielens87}. In our Galaxy, they contain $\\simeq15-20\\%$ of the depleted carbon \\citep[with solar abundance constraints]{zubko04}. As a consequence, they are believed to be part of the interstellar carbon condensation chain \\citep{cherchneff00,dartois05}. From an extragalactic point of view, the luminosity of the $6.2\\mic$ feature can be used as a tracer of star formation \\citep{peeters04}. However, this tracer is biased by global parameters such as the ISM metallicity. Indeed, PAHs are underabundant in low-metallicity galaxies \\citep[{e.g.}][]{galliano03,galliano05,galliano08a,draine07b}. There is a general correlation between the PAH-to-continuum intensity ratio and the ISM metallicity \\citep{madden06,wu06,ohalloran06,smith07}, and consequently between the \\IRACiv/\\MIPSi\\ broadband ratio and the metallicity \\citep{engelbracht05}. The origin of this trend has been attributed to radiative and mechanical destruction mechanisms by \\citet{madden06} and \\citet{ohalloran06} respectively. Conversely, from the detailed modeling of the spectral energy distribution (SED) of nearby galaxies, \\citet{galliano08a} showed that the PAH-to-gas mass ratio at different metallicities coincides with the relative amount of carbon dust condensed in the envelopes of low-mass stars, during the Asymptotic Giant Branch phase (AGB). This study suggests that PAHs are injected into the ISM by their progenitors, the AGB stars, several hundreds of Myr after the beginning of the star formation, when the gas has already been enriched by more massive stars. This delay corresponds to the time needed for AGB stars to evolve off the main sequence. Therefore, the delayed injection of AGB-condensed carbon dust into the ISM offers a natural explanation for the paucity of PAHs in low-metallicity environments. From a cosmological point of view, the large luminosity in these IR emission features coupled with the high sensitivity of \\spitz\\ has allowed the detection of PAHs in distant luminous infrared galaxies out to redshift $z\\simeq2$ \\cite[{e.g.}][]{elbaz05,yan05,houck05}. Hence, understanding what controls the properties of the aromatic features on large scales is required, in order to properly interpret broadband surveys. The detailed characteristics of the mid-IR features, such as their shape, their central wavelength or the intensity ratio between the different bands are known to vary \\citep[see][]{peeters04b}. Such variations are essentially due to modifications of the molecular structure of the PAHs, in different astrophysical environments. In particular, since each feature is attributed to a given vibrational mode, the ratio between these features will vary with quantities such as the charge, the hydrogenation or the size and shape of the molecule. The $3.3\\mic$ PAH band arises from the radiative relaxation of CH stretching modes, while the $11.3\\mic$ feature originates in the CH out-of-plane bending modes; CC stretching modes are responsible for the features between 6 and $9\\mic$; CH in plane bending excitation produces part of the $8.6\\mic$ band. Now, laboratory studies and quantum calculations shows that the CC modes are instrinsically weak in neutral PAHs, and become stronger when the PAHs are ionized \\citep{langhoff96,allamandola99,bauschlicher02,kim02}. Therefore, the 6 to $9\\mic$ bands will be much more intense for a PAH$^+$ than for a PAH$^0$, while it will be the opposite for the 3.3 and $11.3\\mic$. Consequently, the ratios between the CC and the CH feature intensities depend on the charge of the PAHs, which is directly related to the physical conditions (e.g.\\ intensity of the ionizing radiation field, electron density, etc.) in the environment where the emission is originating. Evidence of variations between features in different astrophysical environments have been reported by many authors. For example, \\citet{joblin96} showed that the $\\ipah{8.6}/\\ipah{11.3}$ ratio decreases with increasing distance from the exciting star, in the reflection nebulae \\ngc{1333} --~where $\\ipah{\\lambda}$ is the integrated intensity of the feature centered at $\\lambda\\mic$. \\citet{hony01} found a good correlation between the 3.3 and $11.3\\mic$ CH bands, in a sample of Galactic \\hii\\ regions, YSOs, and evolved stars, while they reported variations of $\\ipah{6.2}/\\ipah{11.3}$ by a factor of 5. The observations of Galactic and Magellanic \\hii\\ regions, presented by \\citet{vermeij02}, indicate that the ratios $\\ipah{6.2}/\\ipah{11.3}$, $\\ipah{7.7}/\\ipah{11.3}$ and $\\ipah{8.6}/\\ipah{11.3}$ are correlated. Furthermore, they suggest a segregation between the values of these ratios in the Milky Way and those in the Magellanic Clouds. \\citet{bregman05} studied the variation of $\\ipah{7.7}/\\ipah{11.3}$ in three reflection nebulae. Assuming that this variation is controlled by the charge of the PAHs, they could relate this band ratio to the ratio $G_0/n_e$ between the integrated intensity of the UV field, $G_0$, and the electron density, $n_e$. Similarly, \\citet{compiegne07}, studying the detailed variations of the mid-IR spectrum in the Horsehead nebula, attributed the high relative strength of the $\\ipah{11.3}$ feature to a high fraction of neutral PAHs, due to the high ambient electron density. On the contrary, \\citet{smith07} studied the variation of $\\ipah{7.7}/\\ipah{11.3}$ coming from the nuclear regions of the SINGS legacy program galaxies. They find that this ratio is relatively constant among pure starbursts, but varies by a factor of 5 among galaxies having a weak AGN. They interpret this effect as a selective destruction of the smallest PAHs by the hard radiation arising from the accretion disk, ruling out the explanation in terms of ionization of the molecules, in these particular environments. This interpretation is also supported by the high $11.3\\mic$ and the weakness of the $3.3\\mic$ band in the \\akari\\ spectrum of the giant elliptical galaxy \\ngc{1316} \\citep{kaneda07}. The previous considerations stress the diversity of the possible interpretation of the mid-IR feature variations in galaxies. We need to identify the main physical processes controling the PAH bands, if we are to use them as diagnostic tools. This is the aim of this paper. It presents a quantitative analysis of mid-IR spectra (\\iso\\ and \\spitz) of Galactic regions, low-metallicity dwarf galaxies, quiescent spirals, starbursts and AGNs. To achieve our goal, we focus on identifying the main trends between the various PAH features, at different spatial scales and in different environments. We then link these variations to the physical conditions inside the studied region. Preliminary results of this study were published in \\citet{galliano04,galliano07c}. The paper is organized as follows. In \\refS{sec:obs}, we present our sample and the data reduction. We discuss the spectral decomposition of the mid-IR spectrum, in \\refS{sec:decomp}. The results of this decomposition are presented in \\refS{sec:correl}; we study the various trends between the band ratios within galaxies, and among different types of environment. Then, in \\refS{sec:explanation}, we provide a physical interpretation of these trends, when the structure of the ISM is resolved, and when it is not. Finally, we summarize our conclusions in \\refS{sec:concl}. ", "conclusions": "\\label{sec:concl} We presented the results of a systematic study aimed at understanding the main properties of the mid-IR features in different astrophysical environments. It is based on observations of Galactic regions and galaxies of various types, observed by the satellites \\iso\\ and \\spitz. We have developed two distinct methods of spectral decomposition with different hypotheses, in order to test the robustness of our trends, and overcome eventual biases. These two methods have shown similar trends between the physical quantities that we have studied. We explored the variations of the different mid-IR features between the wavelengths 5 and $16\\mic$, among integrated spectra of galaxies, and inside galaxies and Galactic regions. Our main results are the followings. \\begin{enumerate} \\item We find that the 6.2, 7.7, $8.6\\mic$ features are essentially tied together, while the ratio of these bands to the $11.3\\mic$ feature can vary by one order of magnitude, in our sample. These variations are seen both inside individual sources (like \\M{82}, \\M{17}, the \\orb, \\xxxdor, etc.), as well as among integrated spectra. In general, the $\\ipah{6.2}/\\ipah{11.3}$ ratio is spatially correlated with the power radiated by the PAHs. It indicates that the ratio $\\ipah{6.2}/\\ipah{11.3}$ is higher in regions of intense star formation. \\item With the help of a stochastic heating model and realistic absorption efficiencies, we show that the variations of the mid-IR features are essentially due to the variation of the fraction of ionized PAHs. We conclude that the properties of the PAHs, throughout our sample, are remarkably universal. In particular, we rule out both the modification of the grain size distribution, and the extinction by the $9.7\\mic$ silicate feature, as an explanation of these variations. Indeed, we show that a modification of the size distribution could explain the observed variation of the $\\ipah{6.2}/\\ipah{11.3}$ ratio. However, it would cause the $\\ipah{6.2}/\\ipah{7.7}$ and $\\ipah{6.2}/\\ipah{8.6}$ (and $\\ipah{3.3}/\\ipah{11.3}$) ratios to vary significantly and to be correlated with the $\\ipah{6.2}/\\ipah{11.3}$, contradicting our observations. Similarly, a deep absorption by the silicate feature would vary the $\\ipah{6.2}/\\ipah{11.3}$ ratio, but it would decouple the $8.6\\mic$ from the 6.2 and $7.7\\mic$ features at the same time. \\item The universality of the properties of the PAHs and the fact that the band ratios are mainly sensitive to the charge of the molecules allow us to use these band ratios as a tracer of the physical conditions inside the emitting region. Using a few well-studied Galactic regions (including the spectral image of the \\orb), we give an empirical relation between $\\ipah{6.2}/\\ipah{11.3}$ and the ratio $G_0/n_e\\times\\sqrt{T_\\sms{gas}}$. \\item In the case where several regions with different physical conditions are integrated, the band ratios are dependent on the morphology of the ISM, as well as on the evolutionary properties of the illuminating stellar cluster. Thus, we find that the $\\ipah{6.2}/\\ipah{11.3}$ band ratios in the star forming nuclei of \\M{51} and \\M{82} are very similar to those in the galactic compact \\hii\\ regions, \\IR{15384} and \\IR{18317}, while the halo of \\M{82} and the spiral arms of \\M{51} are similar to those in the Galactic reflection nebula \\ngc{2023}. These differences in band ratio reflect differences in the ionization over recombination rate and hence trace back to variations in the ratio of the ionizing radiation field to the electron density. \\end{enumerate} \\appendix" }, "0801/0801.1431_arXiv.txt": { "abstract": "Frequencies, powers and damping rates of the solar p modes are all observed to vary over the 11-yr solar activity cycle. Here, we show that simultaneous variations of these parameters give rise to a subtle cross-talk effect, which we call the ``devil in the detail'', that biases p-mode frequencies estimated from analysis of long power frequency spectra. We also show that the resonant peaks observed in the power frequency spectra show small distortions due to the effect. Most of our paper is devoted to a study of the effect for Sun-as-a-star observations of the low-$l$ p modes. We show that for these data the significance of the effect is marginal. We also touch briefly on the likely $l$ dependence of the effect, and discuss the implications of these results for solar structure inversions. ", "introduction": "\\label{sec:intro} Long timebase observations of the ``Sun as a star'' made, for example, by the ground-based BiSON network (Chaplin et al. 2007) and the GOLF (Gabriel et al. 1995) and VIRGO/SPM (Fr\\\"ohlich et al. 1997) instruments on the \\emph{ESA/NASA} SOHO spacecraft, provide key data on the low-degree (low-$l$) p modes for probing the deep solar interior and core. This paper is concerned chiefly with aspects of analysis of these Sun-as-a-star data. Accurate and precise p-mode parameter data are a vital prerequisite for making robust inference on the interior structure of the Sun, and improved precision, and usually also improved accuracy, accrue from use of long helioseismic datasets (e.g., see discussion in Chaplin \\& Basu 2007). However, long helioseismic datasets must by necessity span a sizable fraction, or more, of an 11-yr solar activity cycle. Mode frequencies, damping rates, powers and peak asymmetries (Jim\\'enez-Reyes et al 2007) all vary systematically with the solar cycle, and these variations can give rise to complications for the analysis and interpretation, with implications for parameter accuracies, when long datasets are analyzed in one piece. There is the concern that changes in frequency over a long dataset may distort the underlying shapes of the mode peaks to such an extent that the shapes no longer match the Lorentzian-like functions that are used to model them. This `distortion' effect turns out not to be a significant cause for concern, at least where the Sun is concerned (Chaplin et al., 2008), although it may be an issue for stars that have shorter, and more pronounced, activity cycles than the Sun. There is, however, another potential problem for the solar observations, a `cross-talk' effect that has its origins in the simultaneous variations of the mode frequencies, powers and linewidths over the solar cycle. This cross-talk effect can bias the estimated frequencies of the modes. The effect, which we call the ``devil-in-the-detail'', is the subject of this paper. To introduce the effect, let us begin with a fictitious scenario, in which the p-mode frequencies are the only mode parameters that are observed to vary over the solar cycle. A mode peak observed in the power frequency spectrum of a long dataset will then have a centroid frequency that corresponds to the unweighted average of its time-varying frequency, over the period of observation. This simple rule of thumb holds because the frequency shifts of the low-$l$ solar p modes are small, both in fractional terms and compared to the peak linewidths. In sum: the unweighted time-averaged frequency may be directly recovered. Now consider what happens if the power of the mode is also observed to vary in time. This will mean that timeseries data at different epochs will have varying contributions to the final peak profile observed in the power frequency spectrum. Data samples from times when the mode is more prominent (i.e., at high power) will carry proportionately larger weights in determining the final peak profile than samples from times when the mode is less prominent (i.e., at low power). Because the frequency also varies in time, one would expect the mode peak to be pulled, or biased, to a lower or higher frequency than the unweighted average frequency, with the size and sign of the effect dependent on the comparative time variations of the frequency and power. This is our devil in the detail. Why might the devil in the detail matter? First, it raises uncertainty over what it is we are actually measuring when we determine mode frequencies from long datasets. If we have two sets of data that are not contemporaneous, it creates uncertainty if we wish to correct their frequencies to a common level of activity. Correction procedures, which attempt to `remove' the solar-cycle effects from the fitted frequencies, can be quite sophisticated (e.g., the BiSON correction procedure; see Basu et al. 2007); but they assume that the observed mode frequencies correspond to unweighted time averages of the frequencies. The devil-in-the-detail effect means this is not true in practice. Furthermore, because the frequency bias that is introduced depends on the relative sizes of the parameter variations -- the larger the fractional power variation, for a given frequency variation, the more pronounced the effect -- we might also expect there to be some dependence of the bias on the angular degree, $l$. Both issues above have potential implications for structure inversions made with the estimated p-mode frequencies. The layout of the rest of the paper is as follows. In Section~\\ref{sec:prob}, we spell out why cross-talk from simultaneous variations of the mode parameters introduces the devil-in-the-detail bias in estimation of frequencies from Sun-as-a-star data. We then describe how artificial data were used to illustrate, and quantify, the problem. We give some important background on the Sun-as-a-star observations in Section~\\ref{sec:sas}, and details on the artificial Sun-as-a-star data in Section~\\ref{sec:data}. We then discuss results from analysis of long (9.5-yr) artificial datasets in Section~\\ref{sec:full}, where we attempt to quantify the likely size of the devil-in-the-detail effect for real observations. Finally, we pull together the main points of the paper in Section~\\ref{sec:conc}, where we discuss the implications for analysis of real data. We also touch on the issue of the $l$ dependence of the effect, and the possible impact on results of structure inversions. ", "conclusions": "\\label{sec:conc} Over the course of the solar cycle, variations are observed not only in the p-mode frequencies, but also in the p-mode powers. Variation of the powers over time gives rise to a cross-talk effect with the varying frequencies, which we call the devil-in-the-detail effect. The effect, and its impact, may be summarized as follows. In the absence of any power variation, the location in the power frequency spectrum of the centroid of a mode peak would in principle give an accurate measure of the unweighted average mode frequency over the period of observation. However, when significant time variation of the power is present, data samples from times when the mode is more prominent (i.e., at high power) will carry proportionately larger weights in determining the final peak profile than samples from times when the mode is less prominent (i.e., at low power). As the level of solar activity rises, the frequencies of the low-$l$ p modes under study here increase, but corresponding mode powers decrease. This means that, for a long timeseries, the frequency of a mode peak will be biased so it is \\emph{lower} than the unweighted time-averaged frequency. This is our so-called devil in the detail. \\subsection{Impact on Sun-as-a-star data} \\label{sec:concsas} We have used extensive Monte Carlo simulations to estimate the magnitude of the resulting devil-in-the-detail bias, in BiSON- and GOLF-like Sun-as-a-star helioseismic data spanning almost one full activity cycle. Our simulations imply the bias should rise monotonically in frequency up to $\\approx 3000\\,\\rm \\mu Hz$, and then level off, or even decrease slightly in magnitude, at higher frequencies. It is at approximately the same frequency that the magnitude of the bias reaches a size that is comparable to the expected frequency uncertainties. Should we be worried by this devil-in-the-detail effect? Let us consider briefly, in this and the next section, two issues where the effect might be a potential cause for concern. There is the impact of the effect on attempts to correct the frequencies for solar activity. Since our in-depth analysis has been made for Sun-as-a-star data on the low-$l$ p modes, we are in a position to comment in detail on correction procedures for these data. Procedures which attempt to `remove' the solar-cycle effects from the fitted frequencies (e.g., the BiSON correction procedure; see Basu et al. 2007) assume that the observed mode frequencies correspond to unweighted time averages of the frequencies. The devil-in-the-detail effect therefore introduces a bias in the correction, which will be comparable in size to the frequency uncertainties for modes at frequencies above $\\approx 3000\\,\\rm \\mu Hz$. Will this also be true for analysis of resolved-Sun data on modes of higher $l$? \\subsection{Brief discussion on $l$ dependence of effect} \\label{sec:concl} The issue of the $l$ dependence of the effect is of course important where inversions for the internal structure are concerned, since those inversions must make use of data on modes covering a wide range in $l$. It is a potential complication not just for analysis of non-contemporaneous data from two instruments (or more); but also for analysis of data from a single instrument, since significant variation of the bias with $l$ could affect the accuracy of inversions. The impact of the bias on analysis of medium and high-$l$ data, from resolved-Sun observations, will be covered in detail in a future paper. Here, in order to get a feel for the likely size of the bias, its variation with $l$, and the possible impact on structure inversions, we conducted the following experiment. First, we estimated the size of the devil-in-the-detail bias for different $l$ in the range $0 \\le l \\le 150$, but only at fixed frequency (here, a notional mode frequency of $3000\\,\\rm \\mu Hz$). We assumed the frequency shifts should scale inversely with the mode inertias (and used the model `S' inertias; see Christensen-Dalsgaard et al. 1996); and we estimated the $l$-dependent mode height and linewidth changes from GONG data analyzed in Komm, Howe \\& Hill (2000) [their Figs.~10 and 11). With estimates of the solar-cycle frequency, height and linewidth changes to hand for each $l$, we were in a position to use Equation~\\ref{eq:int} to construct the expected final peak profiles -- assuming a timeseries of length 9.5\\,yr (as in Sections~\\ref{sec:data} and~\\ref{sec:full}) -- and to thereby estimate the devil-in-the-detail bias as a function of $l$. Our results showed little variation of the estimated bias with $l$ below $l \\la 60$ (it being $\\approx 0.02\\,\\rm \\mu Hz$ in size). However, at higher degrees the bias increased in a linear manner with increasing $l$, reaching a size of $\\approx 0.05\\,\\rm \\mu Hz$ at $l \\sim 150$. Next, we scaled the estimated bias for each $l$ by the inertia ratio of Christensen-Dalsgaard \\& Berthomieu (1991), and compared these scaled differences with sizes of scaled frequency differences that are typically large enough to affect inversion results for the internal structure at the $1\\sigma$ level (see also Eff~Darwich et al. 2002). This assumed that the observed p-mode frequencies came from the MDI frequency set of Korzennik (2005), which includes frequencies on modes from $0 \\le l \\le 125$. The results of this comparison suggest that the devil-in-the-detail bias is not quite large enough to affect the inversions at the $1\\sigma$ level, although the amounts by which it fell short were marginal. So, our preliminary conclusion, regarding the impact of the $l$ dependence, is that the devil in the detail may not be a cause for concern where inversions for structure using data from a single instrument are concerned. That said, we should stress that this comparative analysis was an approximate one. At no point did we account for any possible $m$ dependence, which might have a significant influence on the results. Without having done a full Monte Carlo simulation of the data, we were not in a position to provide an accurate estimate of the frequency dependence for different $l$. These factors will be tested in upcoming work. \\subsection*" }, "0801/0801.0810_arXiv.txt": { "abstract": "{We investigate early time inflationary scenarios in an Universe filled with a dilute noncommutative bosonic gas at high temperature. A noncommutative bosonic gas is a gas composed of bosonic scalar field with noncommutative field space on a commutative spacetime. Such noncommutative field theories was recently introduced as a generalization of quantum mechanics on a noncommutative spacetime. As key features of these theories are Lorentz invariance violation and CPT violation. In the present study we use a noncommutative bosonic field theory that besides the noncommutative parameter $\\theta$ shows up a further parameter $\\sigma$. This parameter $\\sigma$ controls the range of the noncommutativity and acts as a regulator for the theory. Both parameters play a key role in the modified dispersion relations of the noncommutative bosonic field, leading to possible striking consequences for phenomenology. In this work we obtain an equation of state $p=\\omega(\\sigma,\\theta;\\beta)\\rho$ for the noncommutative bosonic gas relating pressure $p$ and energy density $\\rho$, in the limit of high temperature. We analyse possible behaviours for this gas parameters $\\sigma$, $\\theta$ and $\\beta$, so that $-1\\leq\\omega<-1/3$, which is the region where the Universe enters an accelerated phase. \\\\ \\vspace{1cm} Keywords: Noncommutative Fields, Lorentz Invariance Violation, CPT Violations, Inflationary Cosmology} ", "introduction": "Quantum Field Theories (QFT) with Lorentz invariance violation are subject of growing attention among theoretical physicists in the last few years \\cite{Kostelecky:1988zi,Kostelecky:1995qk,Colladay:1998fq,Kostelecky:1999mr,Jackiw:1999yp,Kostelecky:2000mm}. Indeed, there is no reason a priori to believe that the cosmological principle\\footnote{The cosmological principle states that spacetime is locally Lorentz invariant.} is true at all energy scales. Further analysis on the problem of matter-antimatter asymmetry \\cite{Dine:2003ax}, ultra high energy cosmic rays \\cite{Anchordoqui:2002hs,Takeda:2002at,Abraham:2006ar}, primordial magnetic field \\cite{Grasso:2000wj}, neutrino physics \\cite{Athanassopoulos:1997pv} and some other cosmological measurements demands further criticisms on the cosmology principle at ultra high energy scale. Therefore the proposal of validity/bound tests for the cosmology principle, i.e., cosmological phenomenology based upon QFT with Lorentz invariance violation is a subject of major relevance. There are distinct approaches to QFT with Lorentz invariance violation. A first approach is based on the noncommutativity of spacetime itself. The most studied situation was that of QFTs on the Groenenwold-Moyal \\cite{Szabo:2001kg,Douglas:2001ba}. Some of these models may be obtained from a fundamental theory such as string/M-theory \\cite{Douglas:2001ba}. Recently \\cite{Chaichian:2004za,Aschieri:2005yw,Aschieri:2005zs}, it was shown that one could restore Lorentz invariance of these QFT by a proper twisting of the action of the symmetry group on the fields. Cosmology applications of this twisted Lorentz invariant QFT is recently being under study \\cite{Akofor:2007fv}. There is also an interesting approach based on using a noncommutative version of Wheeler-deWitt equation, where the noncommutativity is introduced as a Moyal deformation \\cite{GarciaCompean:2001wy}. A second approach is known as extended standard model \\cite{Colladay:1998fq}. It is based on a famous work of S. Carroll, R. Jackiw and G. Field \\cite{Carroll:1989vb}. In this approach the breaking of Lorentz invariance is performed by the introduction of a Chern-Simons term on the action of some gauge theory\\footnote{It is worth mentioning that there is some controversy on the correctness of this approach \\cite{Bonneau:2006ma}}. One should also mention the approach known as doubly special relativity \\cite{Magueijo:2001cr}. Another approach, the one we are going to consider in this paper, is QFT based on noncommutative fields \\cite{Carmona:2002iv,Carmona:2003kh,Balachandran:2007ua}. On such theories one breaks Lorentz invariance by modifying the canonical commutation relations of the fields (therefore the name noncommutative fields) by the inclusion of an ultraviolet or/and infrared scale factor. These noncommutative fields arise as a generalization of the quantum mechanics on the Groenenwold-Moyal plane \\cite{Nair:2000ii}, where one considers the noncommutativity in the degrees of freedom of the fields. One interesting feature of such theories is that there is a connection of this approach with Carroll et al proposal \\cite{Carroll:1989vb} for gauge theories. See for instance \\cite{Carmona:2002iv,Gamboa:2005bf,Falomir:2005it,Gamboa:2005pd,Falomir:2006hp}. One common feature of all these theories with Lorentz invariance broken is the deformation of the dispersion relations of the plane waves. It is this deformation that allows one to build phenomenological models that could be in principle be tested in future cosmology or particle physics experiments. The kind of deformation to be seen or ruled out by these experiments will thus decide which theory will survive. The most promising arena for these phenomenological models is the inflationary scenario of the Universe. Models based on the Universe inflation has successfully explained several observational data, such as the origin of density fluctuations on large scales. The general belief is that inflation that supports large and fast growth of scales can connect the region where the Lorentz invariance might break (i.e., at very high energy) to cosmological scales. On the other hand, it is known that so far there is no explanation of the essential mechanisms of inflation based on fundamental physics such as string/M-theory. In the context of inflationary scenarios based on QFT with Lorentz invariance violation, based on the noncommutativity of space-time, several developments have been put forward recently \\cite{Alexander:2001dr,Lizzi:2002ib,Hassan:2002qk,Koh:2007rx,Koh:2007wa,Chu:2000ww, Bertolami:2002eq}. In these works several efforts have been made in the direction of finding imprints of primordial fluctuations that can survive the inflationary phase. Such imprints have been investigated both in theories with \\cite{Chu:2000ww} and without \\cite{Koh:2007rx} inflaton fields. See \\cite{Brandenberger:2007rg} and references therein for a recent review on this subject. In the present paper, we consider the approach based on noncommutative fields as studied by A. P. Balachandran et al. in \\cite{Balachandran:2007ua}. In that work a noncommutative free massless bosonic field was studied. For that theory, they considered a regularization of the delta function appearing in one of the commutators of the fields. The equal time commutation relation for the fields then reads \\begin{equation} \\label{eq:comm-rel-modified} \\left[ \\hat{\\varphi}^a(x;t),\\hat{\\varphi}^b(y;t)\\right]=i\\epsilon^{ab}\\theta(\\sigma;x-y), \\end{equation} where $\\theta(\\sigma;x-y)$ is a Gaussian distribution with a parameter $\\sigma$ related with the standard deviation of the distribution. Besides introducing one further parameter $\\sigma$, the resulting theory is finite. That fact allowed the study of the deformed black body radiation spectrum. One interesting feature of this deformed spectrum is that the energy density for boson with higher frequency is greater than that for the usual black body radiation spectrum. We investigate here an inflationary model based on an Universe filled with a similar gas composed of a noncommutative boson in the relativistic limit. In this scenario we study the thermodynamics of such gas. We aim at computing and analyzing the equation of state $p=\\omega\\rho$, with $p$ being the pressure, $\\rho$ being the energy density. In the present situation this equation of state is far from trivial. Therefore we have to make some careful approximations, such as considering a power series of the parameters $\\theta$ and temperature $1/\\beta$. Now, for suitable values of the parameters $\\theta$, $\\sigma$ and the temperature $1/\\beta$ of the gas, one obtain $-1\\leq\\omega< -1/3$. For these values of $\\omega$ the Universe presents an accelerated phase. Furthermore, the present case should be contrasted with the tachyon and Chaplygin cosmology that is governed by an equation of state given by the general form $p=-A/\\rho$ \\cite{Gibbons:2003gb}, since here the inflationary regime is followed by a radiation dominated epoch for sufficiently small density $\\rho$. The present paper is organized as follows: in Sec.~\\ref{prel} we review the formulation of the free noncommutative boson field as presented in Balachandran et al. The goal is to obtain the Hamiltonian for such noncommutative field so that we may formulate a statistical quantum mechanics problem for the gas at high temperature; in Sec.~\\ref{stme} we study the gas of noncommutative bosons in the grand canonical formalism. Then by some suitable approximations, we obtain an equation of state $p=\\omega(\\sigma,\\theta;\\beta)\\rho$ for this gas; in Sec.~\\ref{cosmonq} we consider a cosmology model with the noncommutative gas by presenting the proper Friedman equation and then the behavior of the energy density of this gas with the scale parameter. Next we consider, by fine tuning the parameters, possible scenarios yielding inflation; finally in Sec.~\\ref{conclu} we finish the paper with some concluding remarks. ", "conclusions": "\\label{conclu} In this paper we have used noncommutative fields to find cosmological scenarios that can develop inflation at early Universe. A gas composed with a noncommutative bosonic field shows up several interesting properties. In particular, in the dilute regime at high temperature, these properties can be read off from Eq. (\\ref{omega}). The aim of the paper was essentially to find the equation of state of the gas by tackling approximately the statistical quantum mechanics problem for the noncommutative field gas at high temperature. Since we are dealing with highly relativistic bosons not necessarily massless, the approximate relation between energy and momenta $E\\simeq p$ used along the paper is valid. This in turn is consistent with the formulation of free massless boson with noncommutative target space adopted previously. The grand canonical ensemble used in the quantum statistical formulation is also valid because, on the other hand, the masslessness of the bosons is just an approximation. The noncommutative field gas we are dealing with here can develop inflation for a period of time that can be adjusted to observational data. We showed some scenarios that exit inflation phase to a radiation dominated epoch, as it should be according to the perspective of the inflationary cosmology. This feature is certainly a motivation for further studies of these scenarios. In particular, the scenarios with $\\theta$ fixed. In scenarios with $\\sigma$ and $\\beta$ fixed, we are left with only the parameter $\\theta$ free to be determined for a given temperature in Eq. (\\ref{omega_min}). Let us estimate the value of $\\theta$ for $T=10^{14}GeV\\simeq10^{27} K$, the approximate temperature at beginning of the inflation, for $\\omega_{min}\\geq-1$ and $\\omega_{min}<-1/3$, respectively \\ben \\theta_{-1}\\geq 0.4\\times 10^{-25}cm, \\qquad \\theta_{-1/3}<0.5\\times 10^{-25}cm. \\een This result should be contrasted with other astrophysical bounds on the parameter $\\theta.$ For instance, in \\cite{Carmona:2002iv}, it was estimated a value for $\\theta$ to be of the order of $(10^{20} eV)^{-1}\\simeq10^{-25}cm$. They have used the fact that the highest cosmic ray energy be no greater than $\\sim 10^{20} eV$. Recall that the GZK cutoff is of the order of $10^{19} eV$. Furthermore, as we can easily see above, there is a short range for $\\theta$ that allows for inflation in these scenarios. A proper understanding of the nature of the parameter $\\sigma$, for instance, a dynamical theory for it, could shed more light on these developments. Further possibilities may arise if one consider other kinds of regularization for the delta function in (\\ref{eqNCTA:69}). Several other interesting issues such as the implications of the sound speed and the density perturbations can also be investigated to know in what extent the model is in accordance with CMB data. One can estimate the adiabatic sound speed by using \\ben c_s^2=\\left(\\frac{\\partial P}{\\partial \\rho}\\right)_{S,N}=-\\frac{V^2}{m_0N}\\left(\\frac{\\partial P}{\\partial V}\\right)_{S,N}, \\een with $\\rho=m/V=m_0N/V$. Once we have previously assumed a dilute gas $(z\\ll1)$ at high temperature, then by using eqs.(\\ref{therm_quantity2})-(\\ref{log_grand_approx}) one can easily show that the gas satisfies $PV=NT$. The sound speed is then given by \\ben c_s=\\sqrt{\\frac{T}{m_0}}, \\een where $m_0$ is the mass of ultra relativistic bosons satisfying $E^2=m_0^2+p^2\\simeq p^2$. So, at first order in $z$, the approximation maintained throughout this work, the sound speed does not depend on non-commutative parameters. The first correction comes from higher orders in $z$ expansion which can be computed quite easily to be: \\begin{equation} \\frac{T\\,\\left( 1 - \\frac{7\\,z}{64} \\right) }{m_0} + \\frac{91\\,z\\,{\\theta }^2}{16384\\,m_0\\,{\\pi }^2\\,T^3\\,{\\sigma }^6} + \\frac{z\\,\\log (z)}{64\\,m_0\\,T}. \\end{equation} Since in this work we consider the high temperature limit, in which we can disregard the mass, further analysis on the speed of sound cannot be pursued. We let further studies related to this issue for future investigations, together with the low temperature limit." }, "0801/0801.1816.txt": { "abstract": "We have measured the bias of QSOs as a function of QSO luminosity at fixed redshift ($z<1$) by cross-correlating them with Luminous Red Galaxies (LRGs) in the same spatial volume, hence breaking the degeneracy between QSO luminosity and redshift. We use three QSO samples from 2SLAQ, 2QZ and SDSS covering a QSO absolute magnitude range, $-24.5 {\\rm MeV}\\> \\hat =\\> 1.60\\times 10^{-13}\\> {\\rm J} \\> \\hat =\\> 5.08\\times 10^{12}\\> {\\rm m}^{-1}\\> \\hat =\\> 1.52\\times 10^{21}\\> {\\rm s}^{-1}$. In the next section we will discuss the equation of state of an equilibrated ultrarelativistic EPP. Then collective phenomena will be considered. Afterwards transport properties and particle production will be discussed. Finally we will describe properties of an EPP which is not in chemical equilibrium as in the case of laser induced plasmas. We will not consider here the formation process and equilibration of an EPP. Many results presented here can be found in the literature for the case of a QGP (Thoma, 1995a) differing only by numerical factors, e.g. number of degrees of freedom. The purpose of this colloquium is to summarize these results and to extend them to laser induced EPPs as a reference for future experiments. As an example we consider a temperature of 10 MeV as it can be typically realized in laser produced and supernovae EPPs. Laser produced QED plasmas have also been discussed recently in two review articles (Mourou {\\it et al.}, 2006; Marklund and Shukla, 2006) with emphasis on the production mechanism and non-linear effects. Here, however, we want to focus on the properties of an equilibrated EPP as they can be calculated from perturbative QED. Such an EPP in thermal and chemical equlibrium might be the outcome of future laser experiments if the intensity can be increased further on. As an example, we have chosen the predictions of Shen and Meyer-ter-Vehn (2001) based on a numerical simulation (PIC) and cross sections for electron-positron productions, in which two opposite laser beams are focussed on a thin gold foil leading to a chemically non-equiliibrated plasma (see section VI). However, this interesting proposal still waits for its experimental confirmation. Other production mechanisms, based on the Schwinger pair production effect in strong fields (Schwinger, 1951), have shown that pair production can be efficient already far below the Schwinger critical field strength, requiring laser intensities of $5 \\times 10^{29}$ W/cm$^2$, in the case of time dependent and inhomogeneous fields, e.g. two oppositely directed pulsed laser beams in vacuum (see e.g. Avetissian {\\it et al.}, 2002; Narozhny {\\it et al.}, 2004; Di Piazza, 2004; Dunne and Schubert, 2005; Gies and Klingmuller, 2005; Blaschke {\\it et al.} 2006a, Blaschke {\\it et al.}, 2006b, Sch\\\"utzhold {\\it et al.}, 2008). Furthermore the pair production in an X-ray free electron laser (XFEL) has been discussed (Ringwald, 2001; Alkofer {\\it et al.}, 2001; Roberts {\\it et al.}, 2002). QED plasmas in strong magnetic fields have been considered by Danielsson and Grasso (1995) and more general in strong electromagnetic fields by Morozov {\\it et al.} (2002). QED plasmas have also been studied in great detail in the book by Melrose (2008). ", "conclusions": "" }, "0801/0801.4572_arXiv.txt": { "abstract": "Circumstellar disks are an integral part of the star formation process and the sites where planets are formed. Understanding the physical processes that drive their evolution, as disks evolve from optically thick to optically thin, is crucial for our understanding of planet formation. Disks evolve through various processes including accretion onto the star, dust settling and coagulation, dynamical interactions with forming planets, and photo-evaporation. However, the relative importance and timescales of these processes are still poorly understood. In this review, I summarize current models of the different processes that control the evolution of primordial circumstellar disks around low-mass stars (mass $<$ 2 M$_{\\odot}$). I also discuss recent observational developments on circumstellar disk evolution with a focus on new \\emph{Spitzer} results on transition objects. ", "introduction": "It is currently believed that virtually all stars form surrounded by a circumstellar disk, even if this disk is sometimes very short lived. This conclusion follows from simple conservation of angular momentum arguments and is supported by mounting evidence. This evidence ranges from the excess emission, extending from the near-IR (Strom et al. 1989) to the sub-millimeter (Osterloh $\\&$ Beckwith 1995), that is observed in most young (age $<$ 1 Myr) pre-main-sequence (PMS) stars to direct \\emph{Hubble} images of disks (McCaughrean \\& O'Dell 1996) seen as silhouettes in front of the Orion Nebula. Also, even though there is not \\emph{direct} evidence that planets actually grow from circumstellar material, it has become increasingly clear that they are in fact the birthplaces of planets since their masses, sizes, and compositions are consistent with the theoretical minimum-mass solar nebula (Hayashi 1981). Recently, the discovery of exo-planets orbiting nearby main sequence stars has confirmed that the formation of planets is a common process and not a rare phenomenon exclusive to our Solar System. Thus, any theory of planet formation should be robust enough to account for the high incidence of planets and cannot rely on special conditions or on unlikely processes to convert circumstellar dust and gas into planets. Star and planet formation are intimately related. Standard low-mass star formation models (e.g. Shu et al. 1987) describe the free fall collapse of a slowly rotating molecular cloud core followed by the development of a hydrostatic proto-star surrounded by an envelope and a disk of material supported by its residual angular momentum. This early phase is expected to occur on a timescale of about $10^5$ years (Beckwith 1999) and results in an optically revealed classical T Tauri star (CTTS, low-mass PMS star that shows clear evidence for accretion of circumstellar material). This stage is characterized by intense accretion onto the star, strong winds, and bipolar outflows. As the system evolves, presumably into a weak-lined T Tauri star (WTTS, PMS star mostly coeval with CTTSs but that do not show evidence for accretion), accretion ends, and the dust settles into the mid-plane of the disk where the solid particles are believed to stick together and to grow into planetesimals as they collide. Once the objects reach the kilometer scale, gravity increases the collision cross-section of the most massive planetesimals, and runaway accretion occurs (Lissauer 1993). In the standard core accretion model (e.g., Pollack et al. 1996), massive enough proto-planets still embedded in the disk can accrete the remaining gas and become giant planets. The early stages are the most uncertain, and many people suspect that grains will not grow into planetesimals by collisions alone at the rate necessary to go through all stages of giant planet formation before the gas nebula has dissipated. Also, even assuming that grains can grow into macroscopic bodies at the necessary rate, it is still doubtful whether they can form large planetesimals. Once objects reach the meter-size scale, they are expected to rapidly spiral inward due to the dynamical interactions with the gas in the disk, which rotates at a slightly sub-Keplerian velocity because it is supported against gravity by pressure in addition to the centrifugal force. However, since current statistics of extra-solar planets indicate that giant planets are common, the difficulties of the standard core accretion model have led some researchers (e.g., Boss 2000) to revisit an alternative planet formation mechanism that had been put aside for several decades, namely, gravitational instability. In the gravitational instability scenario, giant planets form through the direct gravitational collapse of a massive unstable disk over timescales $<$ 10$^{4}$ yrs. In some hybrid models (e.g., Youdin $\\&$ Shu 2002), solid particles settle to the mid-plane of the circumstellar disk and form a dense gravitationally unstable sub-disk. In this manner, planetesimals form from the gravitational collapse of the material in the mid-plane. The collision of planetesimals leads to runaway accretion, and the process continues in a way analogous to the core accretion model. Thus, while the existence of planets around a significant fraction of all the stars is considered verified, the precise mechanisms through which planets form still remain largely unknown. So far, none of the proposed theories have proven satisfactory. On one hand, the standard model of continuous accretion of solid particles relies on doubtful sticking properties of rocks and on unknown processes to prevent the migration of meter-size objects. On the other hand, gravitational instability relies on unproven mechanisms to enhance the surface density of the disk's mid-plane to trigger the process of planet formation. Clearly, more observational constraints are necessary to help the theoretical work on planet formation to proceed forward. \\begin{figure}[!ht] \\plotone{f1.eps} \\caption{ Inner accretion disk fraction vs. stellar age inferred from H-K excess measurements, binned by cluster or association, for $\\sim$3500 stars from the literature. (Figure taken from Hillenbrand 2006). } \\end{figure} ", "conclusions": "The many transition disks discovered by \\emph{Spitzer} are likely to be the subjects of many follow-up observations and studies for years to come. These are objects that are undergoing rapid transformations and have properties that place them somewhere in between two much better defined evolutionary states: those of regular CTTSs and debris disks. Understanding the physical processes responsible for the diversity of transition disks and its implications for planet formation will be one of main challenges for the field. Another key outstanding disk evolution issue is the question of when the transition from the primordial to the debris disk stage actually occurs. As mentioned in Section 3, some of the WTTS disks have very small fractional luminosities and thus seem to be optically thin. These objects \\emph{could} be younger analogs of the $\\beta$ Pic and AU Mic debris disks, and thus some of the youngest debris disks ever observed. However, this interpretation depends on the assumption that these young WTTS disks are gas poor. It is also possible that some of the WTTS disks have very low fractional disk luminosities as a result of most of their grains growing to sizes $\\gg$10-20 $\\mu$m, in which case they would still be primordial disks. Since a real debris disks requires the presence of second generation of dust produced by the collision of much larger objects, the ages of the youngest debris disks can constrain the time it takes for a disk to form planetesimals. It has even been proposed that detectable second generation dust is not produced until Pluto-sized objects form and trigger a collision cascade (Dominik $\\&$ Decin, 2003). In that case, the presence of debris disks around $\\sim$1-3 Myrs old stars would have even stronger implications for planet formation theories. The fact that, as discussed in Section 3, the incidence of 24 $\\mu$m excesses around low-mass PMS stars with ages $\\sim$10 Myr is $\\sim$0 $\\%$, is consistent with the idea that a quiescent period exists between the dissipation/coagulation of the primordial dust and the onset of the debris phenomenon. However, the confirmation of such a scenario will require far-IR observations sensitive enough to detect, at the distance of the nearest star-forming regions (150-200 pc), debris disks as faint as those observed in the solar neighborhood (10-30 pc). In the near future, \\emph{Herschel} will provide such sensitivity. It will also provide the information on the gas content and grain size distribution of optically thin disks around WTTS required to establish whether they represent the end of the primordial disk phase or the beginning of the debris disk stage. Even though much still remains to be learned about planet formation from the study of circumstellar disks, ultimately, we would like to be able to directly observe planets as they form. ALMA could achieve this milestone in the next decade. With its most extended baseline, ALMA will have a resolution of 0.007$''$ at 625 GHz. This corresponds to an angular resolution of less than 1 AU at the distance of nearby star forming regions like Ophiuchus and Taurus. Narayanan et al. (2006) model the molecular emission line from a gravitationally unstable protoplanetary disk (Figure 5, left panel) and conclude that forming giant planets could be directly imageable using dense gas tracers such as HCO+ (Figure 5, right panel). Such observations could provide the ultimate tests needed to distinguish among the competing theories of planet formation." }, "0801/0801.4091_arXiv.txt": { "abstract": "We have conducted an optical and infrared imaging in the neighbourhoods of 4 triplets of quasars. R, z$^\\prime$, J and K$_s$ images were obtained with MOSAIC II and ISPI at Cerro Tololo Interamerican Observatory. Accurate relative photometry and astrometry were obtained from these images for subsequent use in deriving photometric redshifts. We analyzed the homogeneity and depth of the photometric catalog by comparing with results coming from the literature. The good agreement shows that our magnitudes are reliable to study large scale structure reaching limiting magnitudes of R = 24.5, z$^\\prime$ = 22.5, J = 20.5 and K$_s$ = 19.0. With this catalog we can study the neighbourhoods of the triplets of quasars searching for galaxy overdensities such as groups and galaxy clusters. ", "introduction": "Our present understanding of galaxy formation and evolution can be improved significantly through studies of quasars. In the unified scenario, merger-driven processes include the origin of a quasar by self-regulating, supermassive, black-hole growth \\citep{hop}. Thus, most massive spheroids should host an optically-bright quasar during a brief period of its evolutionary history. There is well-established evidence indicating that several galaxy properties depend on the environment where the galaxies formed and evolved. Several processes - such as stimulated or truncated star formation, tidal stripping, merging, etc. - can determine the nature of galaxy populations and dynamics. Also to be considered is the possibility that AGN feedback processes could induce significant changes in the evolution of galaxies near to the quasar (see for instance \\citet{croton}). By studying quasar environments we may deepen our understanding of the relation of quasar fueling and galaxy interactions and merging; and as a consequence, the formation of bright galaxies. The availability of large quasar samples has increased significantly in recent years, mainly due to the advent of large surveys such as the 2dF Galaxy Redshift Survey \\citep{colles} and the Sloan Digital Sky Survey \\citep{stou}. These samples provide an opportunity for more robust statistical studies of the quasar phenomenon and its link with galaxy formation. At low redshifts (z $\\le $ 0.3), existing cross-correlation analysis of quasar environments show that quasars inhabit environments similar to those of normal galaxies (\\citet{smitha}, \\citet{cold02}. However, on small scales (projected distance, $r_p \\leq 1 \\mpc$; and radial velocity difference, $\\Delta V = 500 \\kms$), the quasar environment is overpopulated by blue disk galaxies having a strong star-formation rate with respect to typical galaxy neighbourhoods \\citep{cold03,cold06}. This result is in agreement with those of \\citet{ilona02,ilona04} in the sense that low-redshift quasars follow the Large Scale Structure traced by galaxy clusters but they are not located near the center of clusters. These quasars are mainly found in the periphery of such structures or between two, possibly merging, clusters. At higher redshifts, quasars are often associated with rich environments \\citep{hall,djor} where the comoving density of galaxies is higher than that expected for the general field. This should be interpreted as the core of future rich clusters. Some results at $z \\sim 1$ suggest that quasars reside on the peripheries of clusters or in cluster mergers \\citep{haines01,haines04,tanaka00,tanaka01}. \\citet{tanaka01}, for example, investigated a group of 5 quasars tracing a $\\sim$ 10 Mpc structure of 4--5 clusters, with only a single radio-loud quasar appearing to be directly associated with one of the clusters. In general, the existing studies of quasar environments suggest their association with forming structures, marking merging clusters and filaments. Consequently, groupings of quasars can be expected to trace regions of extraordinary activity. Pairs of quasars have been strongly studied and they are used to identify cluster of galaxies at high redshifts. \\citet{zhdanov} suggested that quasar pairs can be used as tracers of high redshift large-scale structures. Moreover, \\citet{djor03} found a quasar pair at $z \\sim 5$ associated with large scale structure, possibly a protocluster at that redshift. More recently, \\citet{boris} found that quasar pairs at $z \\sim 1$ are excellent tracers of high density environments suggesting that they can be used to find galaxy clusters. Triplets of quasars in a relative small volume are extremely rare and until now, no cluster-scale triplets have been published or investigated. A very compact quasar association ($\\sim$ 50 Kpc separation) was recently reported by \\citet{djor07} suggesting that this could be a compact galaxy group in a merging process. We propose to use triplets of quasars as ``lighthouses'' to find and study clusters of galaxies that favour such rare events. With this aim in mind, we performed an optical and infrared photometric study of the environments of triplets of quasars at different redshifts on the scale of galaxy clusters. The images allow us to observe galaxies in the early Universe and they are used to investigate the environments of quasar triplets We also study the group and cluster morphology, and the photometric properties of individual galaxies in the neighbourhood of each triplet. We have already found that low-z triplets of Seyfert 1 galaxies are associated with extremely rich clusters of galaxies \\citep{ilona07}. If the quasar triggering mechanism does not evolve strongly over redshift we can expect to find some of the richest galaxy clusters ever discovered at $z \\sim 1$. A negative result will be clear evidence of the evolution of the quasar formation mechanism. This paper is organized as follows: In section 2 we describe the sample selection and section 3 shows the optical and infrared observations and data reduction. Section 4 describes the photometric galaxy catalog, and in section 5 we provide a summary of the main results. We adopt the latest cosmology with $\\Omega_M = 0.3$, $\\Omega_{\\lambda} = 0.7$ and a Hubble constant, $H_0=70 ~ {\\rm{km ~ s^{-1} ~ Mpc^{-1}}}$. ", "conclusions": "We selected a sample of triplets of quasars as those triple systems with percolation longitudes smaller than 2 $\\mpc$ and 2000 $\\kms$. The goal is to use them as ``lighthouses'' to find richest clusters. For this purpose we conducted a photometric study of the neighbourhoods of 4 triplets of quasars in the redshift range of 0.9 to 1.7. Images in R, z$^\\prime$, J and K$_s$ bands were obtained with MOSAIC II and ISPI at CTIO. Magnitudes and coordinates were found to be in excellent agreement with the USNO-A2.0 and 2MASS catalogs. We analyzed the homogeneity and depth of the photometric catalog by comparing the number counts with results coming from the literature. The excellent agreement shows that our magnitudes are reliable, down to limiting magnitudes of R = 24.5, z$^\\prime$ = 22.5, J = 20.5 and K$_s$ = 19.0. These results allow us to study large- scale structure and in a forthcoming paper we will deal with the statistical analysis of this data." }, "0801/0801.1211_arXiv.txt": { "abstract": "We report results based on the monitoring of the BL Lac object Mrk 501 in the optical (B, V and R) passbands from March to May 2000. Observations spread over 12 nights were carried out using 1.2 meter Mount Abu Telescope, India and 61 cm Telescope at Sobaeksan Astronomy Observatory, South Korea. The aim is to study the intra-day variability (IDV), short term variability and color variability in the low state of the source. We have detected flux variation of 0.05 mag in the R-band in time scale of 15 min in one night. In the B and V passbands, we have less data points and it is difficult to infer any IDVs. Short term flux variations are also observed in the V and R bands during the observing run. No significant variation in color (B$-$R) has been detected but (V$-$R) shows variation during the present observing run. Assuming the shortest observed time scale of variability (15 min) to represent the disk instability or pulsation at a distance of 5 Schwarschild radii from the black hole (BH), mass of the central BH is estimated $\\sim$ 1.20 $\\times$ 10$^{8} M_{\\odot}$. {\\bf PACS:} 98.54.Cm, 95.85.Kr, 95.75.De, 95.75.Wx ", "introduction": "Blazars constitute a class of radio-loud active galactic nuclei (AGNs) consisting of BL Lacertae objects (BL Lacs) and flat spectrum radio quasars (FSRQs). In the unified schemes of AGNs, blazars are believed to be the beamed counterparts of the FR I/FR II radio galaxies and many of their properties can be well understood in such a scenario (e.g. Browne 1983; Antonucci 1993; Urry \\& Padovani 1995). According to orientation based unification scheme for radio-loud AGNs, blazar's jet make angles $\\leq$ 20$^{\\circ}$ or so to the line of sight (Urry \\& Padovani 1995). Significant flux variations in the whole range of electromagnetic spectrum on time scale of less than a day to several years are common in blazars. Their radiation at all wavelengths is predominantly non thermal and strongly polarized ($> 3\\%$) in radio to optical bands. Blazar variability can be broadly divided into 3 classes viz. micro variability or intra-night variability or intra-day variability, short term outbursts and long term trends. Variations in flux of a few tenth of a magnitude in a time scale of tens of minutes to a few hours (less than a days) is often known as intra-day variability (IDV) (Wagner \\& Witzel 1995). Short term outbursts can range from weeks to months and long term trends can have time scales of months to several years. Study of optical and near-IR variability of blazars on diverse time scales was carried out by several groups (Miller \\& McGimsey 1978; Xie et al. 1988; Webb et al. 1988; Kidger \\& de Diego 1990; Carini 1990; Carini et al. 1992; Heidt \\& Wagner 1996; Bai et al. 1998, 1999; Fan \\& Lin 1999; Ghosh et al. 2000; Gupta et al. 2002, 2004; Stalin et al. 2005; and references therein). The first convincing evidence of optical IDV was found in BL Lacertae by Miller et al. (1989). The typical optical IDV found in most of the blazars is $\\sim$ 0.01 mag hr$^{-1}$. Carini (1990) observed a sample of 20 blazars and found most of them showing detectable inter-night variations. He also noted that in observing runs of more than 8 hours, the probability of detecting significant IDV was $\\sim$ 80\\%. Subsequently, Heidt \\& Wagner (1996) observed a sample of 34 radio selected BL Lacs and detected IDV in 28 (i.e. 82\\%) sources out of which 75\\% sources showed significant variation with in $<$ 6 hours which further supports the finding of high frequency of occurrence of IDV in blazars. In our recent paper (Gupta \\& Joshi, 2005), we studied the frequency of occurrence of optical IDV in different classes of AGNs and found about 10\\% (18/174) radio-quiet AGNs, 35$-$40\\% (41/115) radio-loud AGNs (excluding blazars) and 70\\% (79/113) blazars to show IDV (in fact $\\sim$ 80$-$85\\% blazars show IDV, if observed for more than 6 hours). Mrk 501 is one of the most interesting and nearby BL Lac object at z = 0.034 which makes it the second closest known BL Lac object after Mrk 421 (z = 0.031). Mrk 501 has been studied for variability in all regions of the electromagnetic spectrum (e.g. Fan \\& Lin 1999, Kataoka et al. 1999, Sambruna et al. 2000, Ghosh et al. 2000, Xue \\& Cui 2005, Gliozzi, et al. 2006, and references therein). In optical region this source has been studied by several groups e.g Stickel et al. (1993) have reported variation of 1.3 magnitude in luminosity; Heidt \\& Wagner (1996) have reported $\\sim$ 32\\% variation in flux in less than two weeks time; Ghosh et al. (2000) have made 10 nights observations on Mrk 501 during March to June, 1997 to search for rapid variability and reported rapid variability in 7 nights; and recently, Fan et al. (2004) observed Mrk 501 for two hours in one night in October 2000 and found no significant variation. The variability in blazars is mainly attributed to either due to the disc instability or the activity in the jet. IDV reported in the flaring and high state is generally attributed to the shock moving down the inhomogeneous medium in the jet. But small amplitude variation in the low state of blazar could be due to instability in the accretion disk (Witta et al. 1991, 1992; Chakrabarti \\& Wiita 1993; Mangalam \\& Wiita 1993). Recently, it is noticed that, in the low luminosity AGNs, accretion disk is radiatively inefficient (Chiaberge et al. 2006; Chiaberge \\& Macchetto 2006; Macchetto \\& Chiaberge 2007; Capetti et al. 2007). So, there will be an alternative way to explain the IDV in the low-state of blazars, in which a weak jet emission will be responsible for the IDV. So far there is no clear answer to this. There is a need to study IDV and short term variability in the optical bands with a good sampling in time when the source is in low state. In view of this we pursued the present study to address both the IDV and short term variability on Mrk 501. This paper is structured as follows: section 2 describes photometric observations and data reductions, section 3 present the results, and discussions and conclusions are presented in section 4. ", "conclusions": "" }, "0801/0801.3214_arXiv.txt": { "abstract": "Famous for the extraordinary richness of its molecular content, the Sgr B2 molecular cloud complex is the prime target in the long-standing search for ever more complex species. We have completed a molecular line survey of the hot dense cores Sgr B2(N) and SgrB2(M) in the 3~mm wavelength range with the IRAM 30\\,m telescope. We performed the analysis of this huge data set by modeling the whole spectrum at once in the LTE approximation. Ongoing analyses yield an average line density of about 100 features/GHz above 3$\\sigma$ for Sgr B2(N), emitted and/or absorbed by a total of 51 molecular species. We find lines from 60 rare isotopologues and from 41 vibrationally excited states in addition to the main species, vibrational ground state lines. For Sgr B2(M), we find about 25 features/GHz above 3$\\sigma$, from 41 molecular species plus 50 isotopologues and 20 vibrationally excited states. Thanks to the constant updates to the Cologne Database for Molecular Spectroscopy, we are working our way through the assignment of the unidentified features, currently 40$\\%$ and 50$\\%$ above 3$\\sigma$ for Sgr B2(N) and SgrB2(M), respectively. ", "introduction": "\\label{s:intro} With several active regions and a total mass of more than $10^6$ M$_\\odot$, Sagittarius B2 (hereafter SgrB2) is one of the most complex and massive sites of star formation in the Galaxy. It is located close to the galactic center and harbors two hot dense cores (M and N), which fragment into several sub cores. SgrB2(N) has a very rich chemistry and was called the Large Molecule Heimat (LMH) by \\citet{Snyder94}. Many complex molecules were discovered there like, e.g., acetic acid \\citep[CH$_3$COOH,][]{Mehringer97}, glycolaldehyde \\citep[CH$_2$(OH)CHO,][]{Hollis00}, and acetamide \\citep[CH$_3$CONH$_2$,][]{Hollis06}. It is therefore a prime target to look for new complex molecules. Line surveys at (sub)mm wavelengths are needed to search for large complex molecules since these molecules emit hundreds of (weak) lines. Regions with rich chemistry produce very crowded spectra with many blended lines and the confusion limit is easily reached. In this context, the secure identification of a new molecule requires the identification of many lines and no single line predicted by a temperature and column density derived from multiple lines should be missing \\citep[see, e.g., the rebuttal of glycine by][]{Snyder05}. Since the spectra are complex, modeling the emission of all known molecules is necessary to prevent mis-assignments and point out blended lines before claiming the detection of a new molecule. To search for new complex molecules, we carried out a complete line survey at 3~mm toward the two hot cores SgrB2(N) and SgrB2(M) and we present here some preliminary results. \\begin{figure}[t] \\begin{center} \\includegraphics[width=9.0cm,angle=270]{belloche_fig1.eps} \\caption{Spectra obtained toward SgrB2(N) (\\textit{top}) and SgrB2(M) (\\textit{bottom}) with the IRAM 30\\,m telescope between 80 and 116 GHz. The blow-ups reveal the full complexity of the SgrB2(N) spectrum and the prominent multi-component absorption lines in the SgrB2(M) spectrum .} \\label{f:survey} \\end{center} \\end{figure} ", "conclusions": "" }, "0801/0801.4328_arXiv.txt": { "abstract": "{} {We searched for massive stars with Balmer-emission consistent with magnetically confined circumstellar material.} {Archival spectroscopic and photometric data were investigated.} {\\object{HR\\,7355} is a formerly unknown He-strong star showing Balmer emission. At $V=6.02$\\,mag, it is one of the brightest objects simultaneously showing anomalous helium absorption and hydrogen emission. Among similar objects, only $\\sigma$\\,Ori~E has so far been subjected to any systematic analysis of the circumstellar material responsible for the emission. We argue that the double-wave photometric period of 0.52\\,d corresponds to the rotation period. In tandem with the high projected equatorial velocity, $v \\sin i=320$\\,km\\,s$^{-1}$, this short period suggests that HR\\,7355 is the most rapidly rotating He-strong star known to date; {a class that was hitherto expected to host stars with slow to moderate rotation only.}} {} ", "introduction": "In the early B-type spectral range a subclass of He-strong stars is found, i.e.\\ stars showing Helium lines with abnormally large equivalent widths. The chemical surface abundances of these stars are influenced by the presence of a strong magnetic field, resulting in a He overabundance that typically varies in strength over the stellar surface. Because He-strong stars are sufficiently luminous to harbour radiatively driven winds \\citep[as diagnosed by ultraviolet absorption line diagnostics; see][]{1990ApJ...365..665S}, they represent ideal laboratories for understanding the process of magnetic wind confinement \\citep[][]{1997A&A...323..121B}. Typically, the fields of these stars are too strong for them to be amenable to the magnetohydrodynamical (MHD) simulations \\citep[e,g,][]{2002ApJ...576..413U}. However, an alternative Rigidly Rotating Magnetosphere (RRM) model for the circumstellar distribution of magnetocentrifugally confined wind plasma by \\citet{2005MNRAS.357..251T} has shown much promise in reproducing the detailed optical variability of the archetype emission-line He-strong star \\object{$\\sigma$\\,Ori~E}. {To date, our knowledge on He-strong stars is limited to slow to moderate rotators, as no rapid rotators had been found. This goes as far as to the conclusion that slow rotation is an intrinsic property of He-strong stars \\citep{1983ApJ...268..195W,1999A&A...345..244Z}, that has to be taken into account by the search for the origin of the magnetic field. This work not only reports the discovery of one more bright massive star to host a magnetosphere for application of the above model, but also extends the parameter range in which He-strong stars are found by a factor of about two in rotational velocity space.} \\object{HR\\,7355} (\\object{HD\\,182\\,180}, \\object{HIP\\,95\\,408}) is a little-observed B2Vn star of $6^{\\rm th}$ magnitude ($V=6.02, B=5.91$), located toward the galactic center. It was listed as a MK-standard by \\citet{1969ApJ...157..313H}, but never examined in detail. Other investigators have instead classified it as B5IV \\citep[see][]{1964PLPla..28....1J}. From studies of larger samples of stars that included HR\\,7355, we know that the star is very rapidly rotating: \\citet{2002ApJ...573..359A} measured $v\\sin i = 320$\\,km\\,s$^{-1}$, while \\citet{2000AcA....50..509G} report $v\\sin i = 270\\pm30$\\,km\\,s$^{-1}$. Hipparcos photometric data indicate that the star is a periodic variable, with $P=0.26\\,$d \\citep{2002MNRAS.331...45K}. \\begin{figure*}[t] \\parbox{18cm}{\\centerline{\\psfig{figure=spectra.ps,width=16cm,clip=t,angle=0}}}% \\caption[xx]{\\label{spectra}Changes in several representative lines between the spectra taken in 1999 (solid line) and 2004 (dotted). The 1999 profile has H$\\alpha$ emission extending out to several times beyond $v \\sin i$ (the latter indicated by the vertical dotted lines, upper left) Note that for the Balmer lines (left column) a wider range in velocity is shown than in the other panels.} \\end{figure*} ", "conclusions": "HR\\,7355 is a previously unknown spectroscopically variable star, and as such it should no longer be used as a spectral standard star. In its capacity as the newest member of the He-strong class, it is not only one of the brighter stars in this class, but is also the most rapidly rotating. In addition to its spectral variability, HR\\,7355 is periodically variable in photometry, with either a single-wave sinusoidal lightcurve of $P_{\\rm sin}=0.260714\\pm0.000003$\\,d or a double-wave pattern with $P_{\\rm dw}=0.521428\\pm0.000006$\\,d. At this point we cannot firmly exclude the possibility that the photometric variations originate in some mechanism other than the surface abundance inhomogeneities. However, the spectra do not show the typical signature of pulsation, and moreover we do not find any other signal in the photometric data consistent with the rotation period. Thus, if the spectral and photometric variations repeat on the same period, then that period is the double-wave period, which also is the rotational period. {It is the first rapidly rotating He-strong star, and as such may pose a challenge to field origin hypotheses that would have led to strong magnetic braking.} We intend to begin a monitoring campaign on the star, to obtain a spectroscopic time series {for further analysis of the fundamental parameters of HR\\,7355, as well as for application of} the framework of the RRM model of \\citet{2005MNRAS.357..251T}. This model has proven extremely successful in explaining the emission-line variations of $\\sigma$\\,Ori~E \\citep{2005ApJ...630L..81T}, and we are optimistic that it can explain the behaviour of HR\\,7355 too." }, "0801/0801.3028_arXiv.txt": { "abstract": "% Puzzles often give birth to the great discoveries, the false discoveries sometimes stimulate the exiting ideas in theoretical physics. The historical examples of both are described in Introduction and in section ``Cosmological Puzzles''. From existing puzzles most attention is given to Ultra High Energy Cosmic Ray (UHECR) puzzle and to cosmological constant problem. The 40-years old UHECR problem consisted in absence of the sharp steepening in spectrum of extragalactic cosmic rays caused by interaction with CMB radiation. This steepening is known as Greisen-Zatsepin-Kuzmin (GZK) cutoff. It is demonstrated here that the features of interaction of cosmic ray protons with CMB are seen now in the spectrum in the form of the dip and beginning of the GZK cutoff. The most serious cosmological problem is caused by large vacuum energy of the known elementary-particle fields which exceeds at least by 45 orders of magnitude the cosmological vacuum energy. The various ideas put forward to solve this problem during last 40 years, have weaknesses and cannot be accepted as the final solution of this puzzle. The anthropic approach is discussed. ", "introduction": "\\begin{center} ALL GREAT DISCOVERIES IN ASTROPHYSICS\\\\ APPEARED UNPREDICTABLY AS PUZZLES.\\\\ WHAT WAS PREDICTED WAS NOT FOUND.\\\\ \\end{center} \\vspace{5mm} Not many good things fall down on us from the sky, but discoveries do. They arrive as the puzzles, and usually disappear as the errors. But sometimes they become real, leaving behind the great discoveries. I will give below a short list of astrophysical discoveries of the last four decades, separating intuitively astrophysics from cosmology.\\\\*[3mm] {\\it Quasars} were discovered in early 1960s as compact radio sources. G. Mathews and A. Sandage in 1960 identified radio source 3C48 with a stellar-like object. M. Schmidt in 1963 deciphered the optical spectrum of quasar 3C273 assuming its redshift, $z=0.158$. Surmounting resistance of sceptics, this explanation moved the source to the distance of 630~Mpc and made its luminosity uncomfortably large, $L \\sim 10^{46}$~erg/s. It was a puzzle, and many respectable astrophysicists spent years trying to squeeze the source back into Galaxy. All of them failed, and in the end the puzzling energy release resulted in the discovery of a {\\it black hole}, an object of general relativity. \\\\ % *[3mm] {\\it Pulsars} were discovered first in 1967 by a student of A. Hewish, Jocelyn Bell. She observed a puzzling periodicity of radio-pulses from an unknown source. After short but intense discussion of different possible sources, including extraterrestrial civilizations and ``little green men'', the magnetized rotating neutron stars, the pulsars, were found to be responsible. It opened a new field of cosmic physics: {\\em relativistic electrodynamics}.\\\\ % *[3mm] {\\it The atmospheric neutrino anomaly and the solar neutrino problem} went along most difficult road to the status of discovery. The puzzling phenomenon in both cases was a neutrino deficit as compared with calculations. The solar neutrino experiments have been started by Ray Davis in 1960 in Brookhaven. With time the solar-neutrino deficit raised, but scepticism of the community raised too. Pushed first by Ray Davis and John Bahcall, the solar neutrino problem moved like a slow coach along a road of three decades long. Fortunately, physics differs from democracy: opinion of majority means usually less than that of ONE. In the end the two obscure puzzles (solar and atmospheric neutrino deficits) have turned into discovery of the most fascinating phenomenon, {\\it neutrino oscillations}, theoretically predicted by Bruno Pontecorvo. \\\\*[3mm] {\\it Supernova SN 1987a} became an elementary-particle laboratory in the sky for a study of properties of neutrinos, axions, majorons {\\it etc}. Detection of neutrinos became a triumph of the theory: the number of detected neutrinos, duration of the neutrino pulse and estimated neutrino luminosity have appeared in agreement with theoretical prediction. Gravitational collapse as a phenomenon providing the SN explosion has been established ... and a puzzle appeared. It was the triumph of the incomplete theory, without rotation of collapsing star. The asymmetric ring around SN 1987a implies that the presupernova was a rotating star (it would be a surprise if not!). Rotation should diminish the neutrino luminosity or even change the collapsing scenario. Why then there is a beautiful agreement between theory and observation? This problem still expects its solution. \\subsection{Greatness of false discoveries} \\begin{center} WHEN FIRST APPEARED THE PUZZLES LOOK WEAK.\\\\ SAVE YOUR TIME SAYING: IT'S RUBBISH.\\\\ IN 90$\\%$ OF CASES YOU WILL BE RIGHT,\\\\ BUT YOU MAY MISS THE GREAT PHYSICS. \\end{center} \\vspace{3mm} False discoveries often have great impact on physics, and {\\it Cyg X-3 } is a famous example.\\\\*[1mm] Cyg X-3 is a galactic binary system well studied in all types of radiations, most notably in X-rays. In 80s many EAS arrays detected from it the 4.8 hour periodic ``gamma-ray'' signal in VHE (Very High Energy, $E\\geq 1$~TeV) and UHE (Ultra High Energy, $E\\geq 0.1-1$~PeV) ranges. The list of these arrays included Kiel, Haverah Park, Fly's Eye, Akeno, Carpet-Baksan, Tien-Shan, Plateau Rosa, Durham, Ooty, Ohya, Gulmarg, Crimea, Dugway, Whipple and others. Probably it is easy to say that there was no single EAS array which claimed no-signal observation. Additionally, some underground detectors (NUSEX, Soudan, MUTRON) marginally observed high energy muon signal from the direction of this source. Apart from the Kiel array, which claimed $6\\sigma$ signal, the confidence level of detection was not high: $3-4 \\sigma$. In 1990 - 1991 two new-generation detectors, CASA-MIA and CYGNUS, put the stringent upper limit to the signal from Cyg X-3, which excluded early observations. Apart from two lessons: \\begin{enumerate} \\item good detectors are better than bad ones, \\item ``$3\\sigma$'' discoveries should not be trusted, even if many detectors confirm them, \\end{enumerate} experience of Cyg X-3 has taught us how to evaluate statistical significance searching for periodic signals. The false discovery of high energy radiation from Cyg X-3 had great impact on theoretical high energy astrophysics, stimulating study of acceleration in binary systems, production of high energy gamma and neutrino radiation and creation of high energy astrophysics with new particles, such as light neutralinos, gluinos {\\it etc}. ", "conclusions": "" }, "0801/0801.4578_arXiv.txt": { "abstract": "We present mid-infrared spectroscopy of a sample of 16 optically faint infrared luminous galaxies obtained with the Infrared Spectrograph (IRS) on the $Spitzer~Space~Telescope$. These sources were jointly selected from $Spitzer$ and $Chandra$ imaging surveys in the NDWFS Bo\\\"otes field and were selected from their bright X-ray fluxes to host luminous AGN. None of the spectra show significant emission from polycyclic aromatic hydrocarbons (PAHs; 6.2$\\rm \\mu m$ equivalent widths $<$0.2$\\rm \\mu m$), consistent with their infrared emission being dominated by AGN. Nine of the X-ray sources show 9.7$\\rm \\mu m$ silicate absorption features. Their redshifts are in the range $0.9$2.6. Instead, they must have only weak silicate absorption or emission features. Comparison with other samples suggests that the fraction of power-law sources appear to increase with X-ray luminosity and decrease with X-ray obscuration. Both the infrared and infrared to optical colors suggest that the sources are relatively unobscured compared with $R$-[24]$>$15 sources. Our X-ray selection is likely to be preferentially biased to less obscured and/or more powerful AGN. We find that sources with silicate absorption features tend to have fainter X-ray fluxes and larger hardness ratios. This suggests that more absorbed sources also have higher column densities of absorbing gas along the same line of sight." }, "0801/0801.4863_arXiv.txt": { "abstract": "We show that as many as $\\sim$50 quasars with at least mJy-level expected flux density can be pre-selected as potential in-beam phase-reference targets for ASTRO-G. Most of them have never been imaged with VLBI. These sources are located around strong, compact calibrator sources that have correlated flux density $>100$~mJy on the longest VLBA baselines at 8.4~GHz. All the targets lie within $12\\arcmin$ from the respective reference source. The basis of this selection is an efficient method to identify potential weak VLBI target quasars simply using optical and low-resolution radio catalogue data. The sample of these dominantly weak sources offers a good opportunity for a statistical study of their radio structure with unprecedented angular resolution at 8.4~GHz. ", "introduction": "Phase-referencing is a way to increase the sensitivity of the Space Very Long Baseline Interferometry (SVLBI) observations that provide extremely high angular resolution due to the baselines exceeding the Earth diameter. Phase-reference imaging in ground-based VLBI is usually done in cycles of interleaving observations between a weak target source and a nearby strong reference source. Delay, delay-rate and phase solutions obtained for the phase-reference calibrator are interpolated and applied for the target source within the atmospheric coherence time, thus increasing the coherent integration time on the weak target source. Unlike the first dedicated SVLBI satellite HALCA \\citep{hira00}, the next-generation satellite ASTRO-G will be capable of rapid attitude changing maneuvers. This, and the accurate orbit determination will allow us to observe suitable nearby reference--target source pairs in the traditional ``nodding'' style \\citep{asak07}. There is another, technically less demanding method which does not require rapid changes in the space antenna pointing if the reference--target separation is so small that both sources are within the primary beam of the 9.3-m ASTRO-G paraboloid antenna ($\\sim12\\arcmin$ at 8.4~GHz). In this scenario, the ground-based part of the SVLBI network performs the usual reference--target switching cycles, while the space antenna remains pointed to the same celestial position. (The diameters of the ground-based VLBI antennas are at least a factor of $\\sim3$ larger, and thus their primary beam sizes are considerably smaller than that of the orbiting antenna.) Successful in-beam phase-referencing experiments have already been conducted with HALCA which could not quickly change its antenna pointing \\citep{pori00,bart00,porc00,guir01}. However, the use of in-beam phase-referencing is severely limited by the small number of sufficiently close source pairs known in the sky. Generally speaking, for any given target source of interest, it is very unlikely to find a suitable phase-reference calibrator within the primary beam of even the relatively small-diameter space antenna. One may reverse the usual logic and select the phase-reference calibrator sources first. {\\it Then} it becomes possible to look for potential weak target sources that are located so close to one of the reference sources that in-beam phase-referencing observations with the orbiting antenna are feasible. Here we show that as many as $\\sim$50 quasars with at least mJy-level expected flux density can be pre-selected as potential in-beam phase-reference targets for ASTRO-G. This prospective sample is large enough for a statistical study of sub-milliarcsecond radio structures of weaker sources in comparison with bright ones. Such a sample, most of which have never been studied with VLBI, would certainly contain individually interesting radio quasars as well. The suitability of at least some of the candidate sources could be verified with ground-based VLBI observations prior to the launch of ASTRO-G. ", "conclusions": "" }, "0801/0801.1488_arXiv.txt": { "abstract": "We present two-dimensional (2D) stellar and gaseous kinematics of the inner $\\sim$\\,130$\\times$180 pc$^2$ of the Narrow Line Seyfert 1 galaxy NGC\\,4051 at a sampling of 4.5\\,pc, from near-infrared $K$-band spectroscopic observations obtained with the Gemini's Near-infrared Integral Field Spectrograph (NIFS) operating with the ALTAIR adaptive optics module. We have used the CO absorption bandheads around 2.3\\,$\\mu$m to obtain the stellar kinematics which show the turnover of the rotation curve at only $\\approx$55\\,pc from the nucleus, revealing a highly concentrated gravitational potential. The stellar velocity dispersion of the bulge is $\\approx$\\,60\\,km\\,s$^{-1}$ -- implying on a nuclear black hole mass of $\\approx\\,10^6$\\,M$_\\odot$ -- within which patches of lower velocity dispersion suggest the presence of regions of more recent star formation. From measurements of the emission-line profiles we have constructed two-dimensional maps for the flux distributitions, line ratios, radial velocities and gas velocity dispersions for the H$_2$, H\\,{\\sc i} and [Ca\\,{\\sc viii}] emitting gas. Each emission line samples a distinct kinematics. The Br$\\gamma$ emission-line shows no rotation as well as no blueshifts or redshifts in excess of 30\\,km\\,s$^{-1}$, and is thus not restricted to the galaxy plane. The [Ca\\,{\\sc viii}] coronal region is compact but resolved, extending over the inner 75\\,pc. It shows the highest blueshifts -- of up to $-250$\\,km\\,s$^{-1}$, and the highest velocity dispersions, interpreted as due to outflows from the active nucleus, supporting an origin close to the nucleus. Subtraction of the stellar velocity field from the gaseous velocity field has allowed us to isolate non-circular motions observed in the H$_2$ emitting gas. The most conspicuous kinematic structures are two nuclear spiral arms -- one observed in blueshift in the far side of the galaxy (to the NE), and the other observed in redshift in the near side of the galaxy (to the SW). We interpret these structures as inflows towards the nucleus, a result similar to those of previous studies in which we have found streaming motions along nuclear spirals in ionized gas using optical IFU observations. We have calculated the mass inflow rate along the nuclear spiral arms, obtaining $\\dot{M}_{H_2} \\approx 4\\times10^{-5}\\,{\\rm M_\\odot\\,yr^{-1}}$, a value $\\sim$\\,100 times smaller than the accretion rate necessary to power the active nucleus. This can be understood as due to the fact that we are only seeing the hot ``skin'' (the H$_2$ emitting gas) of the total mass inflow rate, which is probably dominated by cold molecular gas. From the H$_2$ emission-line ratios we conclude that X-ray heating can account for the observed emission, but the H$_2\\,\\lambda2.1218\\,\\mu$m/Br$\\gamma$ line ratio suggests some contribution from shocks in localized regions close to the compact radio jet. ", "introduction": "The presence of supermassive black holes (SMBHs) at the centres of all galaxies which have stellar bulges is nowadays widely accepted by the astronomy comunity \\citep{gebhardt00,ferrarese00}. According to this scenario the energy emitted by an Active Galactic Nucleus (AGN) is due to the accretion of material onto the SMBH and implies the presence of a gas reservoir close to the AGN. \\citet{lopes07} using archival Hubble Space Telescope (HST) optical images for a large sample of early-type galaxies with and without AGNs, found that all AGN hosts have circumnuclear gas and dust, while this is observed in only 26\\% of a pair-matched sample of inactive galaxies. A gas reservoir close to the AGN is also supported by the presence of recent star formation in the circumnuclear region of active galaxies \\citep{schmitt99,boisson00,storchi-bergmann00,cid-fernandes01,cid-fernandes05,storchi-bergmann05}. However, the strongest signatures that feeding to the SMBH is occurring include the observation of streaming motions in ionized gas along nuclear spirals towards the nucleus of nearby active galaxies using two-dimensional (2D) optical spectroscopy \\citep{fathi06,storchi-bergmann07}. Streaming motions as feeding signatures to the nuclear region have been previously observed in radio wavelengths. \\citet{adler96}, for example, have found streaming motions towards the center along the spiral arms of M\\,81 in H\\,{\\sc i}, while \\citet{mundell99} have found similar streaming motions towards the nucleus along the weak bar of NGC\\,4151. Closer to the center, most of the gas is in the molecular phase, and CO observations have been used to map the gas kinematics and inflows \\citep[e.g.][]{garcia-burillo03,krips05,boone07}. Molecular hydrogen emission lines are also relatively strong in the near-IR $K$-band spectra of active galaxies, and previous studies suggest that its distribution and kinematics is distinct from that observed in the other emission lines, which are usually dominated by outflows \\citep[e.g.][]{crenshaw00,das05}. In the Seyfert galaxies NGC\\,2110 and Circinus, for example, \\citet{storchi-bergmann99} have found broader emission-line profiles for [Fe\\,{\\sc ii}] and Pa$\\beta$ than for the H$_2\\lambda$2.1218$\\mu$m emission-line in a study using near-IR long-slit observations, suggesting a stronger influence of nuclear outflows on the former emission lines and a different origin for the H$_2$ emitting gas, consistent with the colder kinematics of the galaxy disk. More recently, using two-dimensional (hereafter 2D) near-IR spectroscopy of the Seyfert galaxy ESO\\,428-G14, we \\citep{riffel06} have found that the H$_2$ emission distribution was mostly restricted to the plane of the galaxy and was less affected by the AGN outflow than the [Fe\\,{\\sc ii}] and Pa$\\beta$ emission lines. In this paper we use adaptive optics IFU spectroscopic data obtained with the Gemini's Near-infrared Integral Field Spectrograph \\citep[NIFS - ][]{mcgregor03}, in the near-IR $K$-band at a sampling of 0$\\farcs$1$\\times$0$\\farcs$1, to investigate the stellar and gaseous kinematics of the inner $\\sim 3 \\times 4$\\,arcsec$^2$ of the nearby active galaxy NGC\\,4051. The stellar kinematics will be used to constrain the galaxy potential and mass of the SMBH, but our main goal is to look for signatures of feeding mechanisms at parsec scales through the H$_2$ kinematics. NGC\\,4051 is a SABbc galaxy at a distance of only $\\sim$9.3\\,Mpc\\,\\citep{barbosa06}, such that 1$\\farcs$0 corresponds to 45\\,pc at the galaxy. It harbors one of the closest AGN, classified as a Narrow-Line Seyfert 1 (NLS1). We have selected NGC\\,4051 for this study partially on the basis of the recent work of \\citet{barbosa06}, who obtained 2D stellar kinematics using the IFU of the Gemini Multi-Object Spectrograph (GMOS) to observe the stellar absorption lines of the Calcium triplet around 8500\\AA. These authors have shown that the turnover of the stellar rotation curve occurs at only $R\\sim$50\\,pc from the nucleus, indicating that the stellar motions are dominated by a highly concentrated gravitational potential. As NIFS with the ALTAIR adaptive optics module provides a much better image sampling and resolution than the GMOS-IFU, we decided to further investigate the kinematics of the nuclear region of NGC\\,4051 in order to better constrain the gravitational potential and the mass of the SMBH. In adition, this galaxy shows strong H$_2$ emission \\citep{rogerio06}, making it a good candidate for a study of its kinematics. Previous studies of NGC\\,4051 include HST narrow-band [O\\,{\\sc iii}] images which show an unresolved nuclear source and faint extended emission (by 1\\farcs2) along the position angle PA=100$^\\circ$ \\citep{schmitt96}, the approximate direction of the alignment of two radio components at 6\\,cm separated by 0$\\farcs$4 \\citep{ulvestad84}. \\citet{veilleux91} reported that the profiles of optical forbidden emission lines present blue wings reaching velocities of up to 800\\,km\\,s$^{-1}$, and proposed a model for the narrow-line region (hereafter NLR) with outflows and obscuring dust. Evidence for outflows have also been observed by \\citet{christopoulou97} for the [O\\,{\\sc iii}] emission to 1\\farcs5 NE of the nucleus. \\citet{nagao00} using the [Fe\\,{\\sc X}]$\\lambda$6374 emission line report a high ionization region extending to 3$\\farcs$0\\ SE of the nucleus. \\citet{lawrence85} found a X-ray variability on time scales of the order of 1 hour and \\citet{salvati93} reported flux changes by a factor of 2 in 6 months observed at 2.2\\,$\\mu$m and suggest that emission by dust reprocessing of the UV radiation from the nucleus is an acceptable explanation. \\citet{ponti06} modelled the nuclear emission by two power law components, one due to the AGN and other due to reflection by the accretion disk. This paper is organized as follows: in section 2 we describe the observations and data reduction. In section 3 we present the results for the stellar kinematics. In section 4 we present the emission-line flux distributions and in section 5 we present the results for the gas kinematics. In section 6 we discuss the results and in section 7 we present the conclusions of this work. ", "conclusions": "We have analysed two-dimensional near-IR $K-$band spectra from the inner $\\approx$\\,150\\,pc of the Narrow Line Seyfert 1 galaxy NGC\\,4051 obtained with the Gemini NIFS integral field spectrograph at a sampling of 4.5$\\times$4.5\\,pc$^2$ at the galaxy and spectral resolution of $\\approx$3\\,\\AA. We have mapped the stellar and gaseous kinematics, and the emission-line flux distributions and ratios of the molecular and ionized hydrogen. The main conclusions of this work are: \\begin{itemize} \\item The turnover of the stellar rotation curve is at only $\\approx$55\\,pc from the nucleus, suggesting that the stellar motions are dominated by a highly concentrated gravitational potential, a result confirmed by modeling using a Plummer gravitational potential, as we obtain a small scale length, $A\\approx$39\\,pc. This result supports the findings of \\citet{barbosa06} based of optical data. The mean velocity dispersion of the bulge ($\\sim$60\\,km\\,s$^{-1}$) implies a SMBH mass of $\\sim 10^6$\\,M$_\\odot$. Within the bulge, we find patches of lower velocity dispersions, which we attribute to recent star formation. \\item The gas kinematics is distinct for each emission line: the Br$\\gamma$ emission line shows velocities restricted to within $\\sim$30\\,km\\,s$^{-1}$ from systemic, without evidence of bultk motions, suggesting that the ionized gas is not restricted to the galactic plane, while much larger blueshifts and redshifts are observed for both the [Ca\\,{\\sc viii}] and H$_2$ emitting gas. The Br$\\gamma$ velocity dispersion is small ($\\sim$40\\,km\\,s$^{-1}$) over most of the field, except around the nucleus at a location close to the SW tip of the radio jet, suggesting some energy injection in the gas by interaction with the radio jet. \\item The [Ca\\,{\\sc viii}] coronal emission line is compact but resolved, extending over a circular region of radius $\\approx$35\\,pc around the nucleus. It also present the largest velocity blueshifts ($-250\\,$km\\,s$^{-1}$) and velocity dispersion (150\\,km\\,s$^{-1}$) of the emission lines studied here, supporting an origin close to the active nucleus, possibly in the transition region between the BLR and the NLR. \\item The H$_2$ emitting gas seems to be mostly restricted to the galaxy plane. The most conspicuous kinematic features are a curved elongated blueshifted structure to the NE, interpreted as gas inflow along a nuclear spiral arm in the far side of the galaxy, and a curved, redshifted structure to the SW, interpreted as gas inflow along a nuclear spiral arm in the near side of the galaxy. Estimates of the mass inflow rate in hot H$_2$ gives $\\dot{M}_{H_2} \\approx 4\\times10^{-5}\\,{\\rm M_\\odot\\,yr^{-1}}$, which is $\\sim\\,100$ times smaller than the nuclear accretion rate necessary to power the active nucleus of NGC\\,4051, supporting the presence of additional inflow of cold non-emitting gas. This is not the first time we have been able to map inflows towards an active nucleus along nuclear spiral arms. In two previous studies we \\citep{fathi06,storchi-bergmann07} have mapped streaming motions towards the active nucleus along nuclear spirals in the galaxies NGC\\,1097 and NGC\\,6951 although in ionized gas. As nuclear spirals are ubiquitous around active galactic nuclei \\citep{lopes07,prieto05}, such inflows seem to be an ``universal'' mechanism to bring gas inwards to feed the SMBH within the inner few hundred parsecs of the galaxy. \\item The total mass of hot H$_2$ is estimated to be of the order of 66\\,M$_\\odot$, while that of H\\,{\\sc ii} is estimated to be 1.4$\\times\\,10^5$ M$_\\odot$. \\item From the H$_2$ line ratios we conclude that H$_2$ is excited by thermal processes -- heating by X-rays from the AGN and shocks produced by the radio jet. We conclude, based in X-ray excitation models of \\citet{maloney96} that X-ray heating can account for the observed emission, but the H$_2\\,\\lambda2.1218\\,\\mu$m/Br$\\gamma$ line ratio supports some contribution from shocks in the regions where the radio jet interacts with the H$_2$ emitting gas. \\end{itemize}" }, "0801/0801.2674_arXiv.txt": { "abstract": "We first summarize work that has been done on the effects of binaries on theoretical population synthesis of stars and stellar phenomena. Next, we highlight the influence of stellar dynamics in young clusters by discussing a few candidate UFOs (unconventionally formed objects) like intermediate mass black holes, $\\eta$ Car, $\\zeta$ Pup, $\\gamma$$^{2}$ Velorum and WR 140. ", "introduction": "\\medskip Many (most?) of the massive stars are formed in clusters and the observations reveal that many cluster members are binary components (to illustrate, see Sana et al., 2007). The effect of binaries on population studies has been discussed many times by different research groups (the Brussels group was one of them) but the following points may be worth repeating: \\medskip \\begin{itemize} \\item binaries with initial primary mass larger than 40 $M_{\\odot}$ and with a period that is large enough so that LBV mass loss happens before the Roche lobe overflow (RLOF) would start, avoid RLOF (the LBV scenario as it was introduced in Vanbeveren, 1991); this may not be true for case A binaries \\item most of the primaries of binaries with orbital period as large as 10 yrs and with initial mass smaller than 40 $M_{\\odot}$ lose most of their hydrogen rich layers by RLOF; the evolution of these stars is therefore quite different from the evolution of single stars with the same mass \\item the RLOF in late case B and case C binaries may be accompanied by the common envelope process that may result in the merger of the two components; it can be expected that this merger will be a rapid rotator with peculiar chemical abundances \\item the RLOF in Case A and early case B binaries implies mass and angular momentum transfer towards the gainer; together with the merger process discussed above, this process is a natural way to make rapid rotators and therefore massive close binaries may be natural progenitors of long gamma ray bursts (Cantiello et al., 2007) \\item the favored model for short gamma ray bursts is the binary merger of two relativistic compact stars and it cannot be excluded that such mergers are important sites of galactic r-process enrichment (De Donder and Vanbeveren, 2003a) \\item\tthe observed massive binary frequency is not necessarily the binary frequency at birth. This is due to the fact that a significant fraction of all massive binaries become single stars during their evolution (many binaries with small mass ratio merge, late case B and case C binaries evolve through a common envelope phase and may merge as well, the supernova explosion of one of the components disrupts the binary), e.g. many single stars may have a binary past with an evolution which is distinctly different from stars that are born as single stars. \\end{itemize} \\medskip The proceedings of the conference on 'Massive Stars in Interacting Binaries' (eds. A. Moffat and N. St.-Louis) were published a few months ago. They can be considered as an update of the monograph 'The Brightest Binaries' published in 1998 (Vanbeveren et al. 1998c) and we recommend careful reading to anyone with an open mind willing to admit that many theoretical population studies, where the effects of binaries are ignored, have mainly an academic value. In section 2 we list a few extra references that may be interesting in order to learn more about the effects of massive binaries on population studies. Many massive objects (single or binary) are born in dense (embedded) clusters, containing a few 10s up to a few 1000s massive stars and even more (e.g., the clusters in the Carina association, the Arches and Quintuplet clusters in the galactic bulge, the Orion nebula cluster, R136 in the LMC, MGG-11 in M82, etc.). In these clusters the evolution of each object may be affected by the presence of all the others, or at least by the closest neighbors. In sections 3 and 4 we discuss a few peculiar massive objects that may be the result of the dynamical evolution of stars in dense clusters, we like to call them Unconventionally Formed Objects (UFOs). ", "conclusions": "" }, "0801/0801.0671_arXiv.txt": { "abstract": "We provide a new interpretation of ultraviolet transition region emission line widths observed by the SUMER instrument on the Solar and Heliospheric Observatory ({\\em SOHO}). This investigation is prompted by observations of the chromosphere at unprecedented spatial and temporal resolution from the Solar Optical Telescope (SOT) on {\\em Hinode} revealing that all chromospheric structures above the limb display significant transverse (\\alfvenic{}) perturbations. We demonstrate that the magnitude, network sensitivity and apparent center-to-limb isotropy of the measured line widths (formed below 250,000K) can be explained by an observationally constrained forward-model in which the line width is caused by the line-of-sight superposition of longitudinal and \\alfvenic{} motions on the small-scale (spicular) structures that dominate the chromosphere and low transition region. ", "introduction": "\\citet{Alfven1947} speculated that the broadening of coronal emission lines was the signature of \\alfven{} waves of sufficient strength to be a potential heating source for Edl\\'{e}n's recently observed hot solar corona \\citep[][]{Edlen1943}. The physical origin of the enhanced UV (and EUV) transition region (TR) and coronal line widths, and the amount of ``hidden'' energy that they represent, has been an open topic of debate, conflicting interpretation and conjecture in the community since. Interest in \\alfven{} waves in this regard is still strong \\citep[e.g.,][]{Chae1998, Erdelyi1998, Banerjee1998, Moran2003, Peter2003, OShea2005} owing to the recent numerical models of quiet coronal heating and fast solar wind acceleration depend critically on the presence of a significant energy carried by them \\citep[e.g.,][]{Suzuki2005, Cranmer2005, Verdini2007}. Only one thing has prevented the validation of the observational and theoretical inferences; the direct observation of \\alfven{} waves in the solar atmosphere. Recent unequivocal observations of low-frequency ($<$5mHz) propagating \\alfvenic{} motions in the solar corona \\citep[][]{Tomczyk2007} and chromosphere \\citep[][]{DePontieu2007b}, plus the relationship of the latter with finely structured spicules observed at the solar limb \\citep[][]{DePontieu2007a} with the Solar Optical Telescope \\citep[SOT;][]{SOT} on {\\em Hinode} \\citep[][]{Hinode}, have inspired us to look again at the relationship between non-thermal emission line widths and spatially unresolved \\alfvenic{} motions\\footnote{For a discussion on why we call these motions \\alfvenic{} and not MHD kink-mode waves, see the supporting online material of \\citet{DePontieu2007b}}. A reappraisal of TR line widths is clearly necessary. Much of the previous scientific effort in this area, prior to the launch of the Solar and Heliospheric Observatory \\citep[{\\em SOHO};][]{Fleck+others1995}, is expertly summarized in Ch.~5.2 of \\cite{Mariska1992}, but we will focus our discussion on the measurements of the Solar Ultraviolet Measurement of Emitted Radiation \\citep[SUMER;][]{Wilhelm1995} instrument. We present a new analysis of SUMER full-disk spectroheliograms taken in early 1996 \\citep[e.g.,][]{Peter1999a,Peter1999b} paying close attention to the spatial variation of the line intensities and non-thermal widths. We compare the results of the data analysis with those of an observationally constrained forward model based on the relationship of chromospheric flows and \\alfven{} waves that occur on spatial structures that are considerably smaller than the SUMER spatial resolution (1\\arcsec). The synthesis of data and model analysis demonstrates that the superposition of transverse and longitudinal motions on sub-resolution structure can explain the observed magnitude, network sensitivity and center-to-limb isotropy of the measured non-thermal line broadening in cooler ($<$250,000K) TR lines. The last finding leads us to the conclusion that previous arguments made against the presence of \\alfven{} waves in the TR, based on the apparent isotropy of the line width variation \\citep[e.g.,][]{Chae1998}, are invalid. ", "conclusions": "Our results indicate that the SUMER \\ion{C}{4} observations of line widths are indeed fully compatible with the LOS superposition of a large number of finely scaled spicules that carry significant longitudinal mass flows as well as vigorous \\alfven{} waves. Since the longitudinal and transverse velocity components are of roughly the same order, the overall center-to-limb variation is relatively limited, naturally producing an isotropy of line widths which had been used as an argument against \\alfven{} waves in the past \\citep[e.g.][]{Chae1998}. There is a slight but significant increase of line widths in a region close to and off the limb, which seems to be caused by \\alfvenic{} motions along the much increased number of structures (the first few 2\\arcsec{} pixels around the limb contain spicules from three neighboring network elements!). The steep drop-off of the line width as we go above a certain height above the limb is directly related to the dwindling number of spicular structures: spicules have typical heights of 5,000 km with very few above 10,000 km. This straightforward interpretation contradicts earlier work that explains this decrease as a result of ion-cyclotron damping of high-frequency \\alfven{} waves \\citep{Peter2003}. Our model also suggests that the increase of line widths in and around the magnetic network is directly related to the concentration of spicule flows along the field {\\em and} \\alfvenic{} motions. We expect that similar results will hold for UV emission lines formed in the low TR (T$<$250,000K) although we note that opacity plays a significant role in the broadening of cooler lines \\citep[e.g.,][]{Doschek2004}. It is possible that \\ion{C}{4} has some opacity at the limb as well, explaining why the forward model predicts intensity increases at the limb that are twice those observed. % It is interesting that, for emission lines formed at temperatures above 500,000K (e.g., like those of \\ion{Ne}{8}), the bump at the limb does not appear, instead we observe a linear increase of line width with the decreasing density above the limb, indicative of the undamped growth of \\alfven{} waves in the extended corona \\citep[e.g.,][]{Banerjee1998}. Our model does not apply to these hotter lines, since the spatial distribution of emission from these lines is no longer dominated by spicules, but rather by thermal conduction and the overlying coronal structure. Further study will be necessary to explain the larger (by $\\sim$5km/s) line widths in polar and equatorial coronal holes. These enhancements may well be caused by the lower densities and the resulting higher amplitude of the \\alfven{} waves (for an equal energy flux). Detailed studies with SOT are necessary to determine whether different amplitudes for the spicular flows perhaps also play a role, especially since the mix of spicule types changes considerably in coronal holes \\citep{DePontieu2007b}. We note that the absolute magnitude of the line width is of significant importance in constraining the amplitude of the \\alfven{} waves that are needed for solar wind models. Unfortunately, the literature reveals a wide scatter of values for the same TR UV emission lines \\cite [see, e.g.,][]{Mariska1992,Chae1998} from different instruments, epochs, locations, etc. This is perhaps not surprising since our model shows that we can expect significantly different values of line widths for the same line depending on proximity to the limb, the small-scale distribution of magnetic flux (which varies considerably over the solar cycle) and the orientation of the magnetic field (since inclined spicules will show varying contributions of longitudinal and transverse velocities). In addition to these physical causes, we have found that the fitting algorithm and form of functional fit (Gaussian with constant, linear or quadratic background) can have a significant impact ($\\sim 10$ km/s) on the magnitude of the line width; e.g., the more complex the background form of the fit the smaller the resulting line width. Our forward model is only a first, primitive, step towards a comprehensive understanding of what dominates the emission of the TR. There has been extensive discusson on the detailed physical nature of the TR in the past \\citep[e.g.,][]{Judge1999,Peter2000a}, much of which has ignored the intrinsically dynamic nature of the TR \\citep[see, e.g.,][]{Wikstol1998}, which is inevitable given its connection to spicules. It is clear from our results that more sophisticated models of the TR are needed that take into account the intrinsically dynamic and finely structure nature of its dominant components." }, "0801/0801.2168_arXiv.txt": { "abstract": "We summarize three recent publications which suggest that the Galactic center region Sagittarius B (Sgr B) may contain non-thermal radio components (Crocker et al. 2007, Hollis et al. 2007 and Yusef-Zadeh et al. 2007a). Based on new VLA matched-resolution continuum data at 327 MHz and 1.4 GHz, we find no evidence for large scale non-thermal radio emission at these frequencies; the spectral behavior is likely determined by the complex summation of multiple HII region components with a wide range of emission measures and hence radio turn-over frequencies. Also, we discuss a possible additional interpretation of the radio continuum spectrum of individual component Sgr B2-F carried out by Yusef-Zadeh et al; confusion from nearby HII components with widely different turn-over frequencies may contribute to the the change in slope of the radio continuum in this direction at low frequencies. Finally, we discuss the uncertainties in the determination of the spectral index of the GBT continuum data of Sgr B carried out by Hollis et al. We find that the apparent spectral index determined by their procedure is also likely due to a summation over the many diverse thermal components in this direction. ", "introduction": "The radio spectrum of the Galactic Center (GC) has been a longstanding and complicated problem for astronomers. The GC radio source was slowly recognized to be a complex of individual sources embedded in diffuse Galactic background emission. By 1965, when Bernard Burke's review article, entitled \"Radio Radiation from the Galactic Nuclear Environment\", appeared in the {\\it Annual Review of Astronomy and Astrophysics}, images of the GC region had been made at a range of frequencies as high as 14 GHz. One of the images in this review article was from Cooper \\& Price (1964), made with the Parkes, Australia, 210-foot telescope at 3 GHz. This image had a resolution of 6.7\\arcmin, making it possible to resolve the Sagittarius B (Sgr B) complex from the other major GC sources. Cooper \\& Price (1964) were the first to suggest that the Sgr B source is an optically-thick thermal source based on a spectral index measurement between 3 and 8 GHz (Drake 1959a,b) that had large uncertainties. During the last 40 years, the Sgr B region has been the subject of numerous detailed studies carried out with the increasing resolution and sensitivity of new instrumentation. \\subsection{The SgrB Region: Sgr B2 and Sgr B1} The Sgr B region is now understood to be one of the most complex star-forming regions in the Galaxy. Located at $\\sim$105 pc in projection from Sgr A$^*$, the Sgr B giant molecular cloud has a total mass of 8-20 x 10$^6$ \\msun~(Tsuboi et al. 1999). Therefore, many interstellar molecules have been detected for the first time in the Sgr B cloud. Most studies of Sgr B have been carried out in the far-infrared and radio regimes because this region of the Galaxy is highly obscured (by up to 50 magnitudes) at visible wavelengths. The Sgr B region is comprised of two distinct complexes: Sgr B2 (G0.7-0.0; to the North) and Sgr B1 (G0.5-0.0; to the South). Both sources are strong radio continuum sources and both have been well-studied on fine scales ($\\sim$1\\arcsec~or better) with interferometric observations. In higher frequency radio observations ($\\nu$ $>$ 1.4 GHz), Sgr B2 dominates due to the numerous ultra-compact, optically-thick HII regions (Benson \\& Johnston 1984; Gaume \\& Claussen 1990). Sgr B2 is well-known for its incredibly high densities ($>$ 10$^5$ cm$^{-3}$) in both ionized and molecular gas (Huttemeister et al. 1993; DePree, Goss \\& Gaume 1998). The Sgr B2 complex has three bright cores - Sgr B2 ``Main'', ``North'' and ``South''. Sgr B2 shows radio emission over a factor of 6000 in size scale, from $<$1\\arcsec~(ultra-compact HII regions) to the 10\\arcmin~extent of diffuse emission along the complex. The highest resolution radio study of Sgr B2 was made at 43 GHz with the Very Large Array (VLA)\\footnotemark\\footnotetext{The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.} radio telescope with an angular resolution of $\\sim$64 milli-arcseconds (600 AU at the GC distance of 8 kpc) in SgrB2 Main. This study revealed over 20 ultra-compact HII regions and stellar wind sources in Sgr B2 Main alone (De Pree et al. 1998). In addition, the range of observed emission measures in the ionized gas are from 1$\\times$10$^4$ to 1$\\times$10$^9$ pc cm$^{-6}$. Radio recombination line studies at numerous frequencies have helped cement the thermal nature of many of the individual radio components in the Sgr B complex (e.g., Mehringer et al. 1993). The radio emission arising from the Sgr B1 complex indicates that this region is a more highly evolved star-forming region with numerous filamentary and shell-like ionized structures (Mehringer et al. 1992). Lying between Sgr B1 and Sgr B2 is a complex of radio sources known as G0.6-0.0. Gas velocities and morphology suggest that G0.6-0.0 physically connects Sgr B1 and Sgr B2 (Mehringer et al. 1992). \\subsection{Recent Papers: Non-thermal Emission in Sgr B?} Three recent papers on Sgr B complex have addressed the nature of the radio emission arising from this region, and have argued that there is a non-thermal component of the radio emission (Crocker et al. 2007; Hollis et al. 2007; and Yusef-Zadeh et al. 2007a). Because Sgr B is one of the most active regions of star formation in the Galaxy, containing more than 60 HII regions, it would not be surprising to find an embedded supernova remnant or two. However, because the preponderance of evidence for the last four decades has indicated that the Sgr B complex is a thermal radio source, claims of non-thermal emission deserve close scrutiny. One of the main motivations for looking for non-thermal emission has come from high-energy investigations of this region. Using HESS, Aharonian et al. (1996) have made a detection of diffuse $\\gamma$-ray flux between 0.2-20 TeV (1 TeV = 10$^{12}$ eV) from the Galactic center (GC) region, distributed along the Galactic plane. In addition, a separate HESS measurement of the $\\gamma$-ray flux and spectrum in a 0.5\\arcdeg$\\times$0.5\\arcdeg~region was centered on the SgrB molecular cloud. After removal of several bright point-like sources, the diffuse flux is believed to be correlated with the density of molecular gas (H$_2$) in the GC region, which has been imaged in the CS (J=1-0) line transition (Tsuboi et al. 1999). One possible origin of the high-energy radiation is the collision between cosmic ray protons and the ambient, dense molecular gas in the GC. Crocker et al. (2007) predict the radio synchrotron spectrum from their broadband (radio to $\\gamma$-ray) emission models and compare their results to the measured radio spectrum of this region. Crocker et al. (2007) base their radio spectrum of this region primarily on low-frequency data at 327 and 843 MHz (see below for a description). They determine that the radio spectrum of Sgr B is non-thermal in nature, with a measured excess of non-thermal emission. In addition, Yusef-Zadeh et al. (2007a) consider the global heating of molecular clouds by cosmic rays, based on their earlier observations of correlations between molecular gas and 6.4 keV K$\\alpha$ line emission (Yusef-Zadeh et al. 2007b). In order to illustrate that SgrB2 has a non-thermal component, Yusef-Zadeh et al. (2007a) use low frequency observations between 255 MHz and 1.4 GHz to show that a part of the SgrB2 complex (a well-studied part, SgrB2 ``F'', a cluster of ultra-compact HII regions) has a component of non-thermal emission with a flux density of $\\sim$82 mJy. They suggest the presence of enhanced heating by cosmic ray particles (which increases the ionization fraction) in the molecular gas. Finally, Hollis et al. (2007) present a study of the continuum temperature of Sgr B2 based on GBT spectral line observations. They argue that there is a non-thermal component in Sgr B2 on size scales of $\\sim$143\\arcsec~at 1.4 GHz with an optically-thin spectral index of $\\alpha$=$-$0.7. In this paper, a discussion of the measurements used to derive the spectral index of radio emission in the above papers follows. In addition, a high-resolution 1.4 GHz image of the GC region is presented and used for comparison to the three recent results on the non-thermal emission in Sgr B. ", "conclusions": "In this paper, we have summarized three recent papers which point out possible non-thermal radio emission arising from the Sgr B region in the GC. We also present a high-resolution and sensitive VLA image of the Sgr B region at 1.4 GHz. Using this image and a matched-array 327 MHz VLA image, we derive a thermal spectrum for the Sgr B complex and suggest that the radio emission is a mixture of optically thin and optically thick emission over the frequency range discussed here (255 MHz to 1.4 GHz). In addition, we show that the the apparent non-thermal power law slope for the Sgr B2 continuum temperature observed by the GBT is likely determined by source structure and provides limited information about the physical processes in the Sgr B region. While the structure Sgr B region is complex and furthermore confused by the Galactic background, there does not appear to be substantial evidence for a non-thermal component in the Sgr B complex. The authors would like to thank Roland Crocker, Mike Hollis, and Farhad Yusef-Zadeh for their comments on this article." }, "0801/0801.0995_arXiv.txt": { "abstract": "Large-scale structure formation, accretion and merging processes, AGN activity produce cosmological gas shocks. The shocks convert a fraction of the energy of gravitationally accelerated flows to internal energy of the gas. Being the main gas-heating agent, cosmological shocks could amplify magnetic fields and accelerate energetic particles via the multi-fluid plasma relaxation processes. We first discuss the basic properties of standard single-fluid shocks. Cosmological plasma shocks are expected to be collisionless. We then review the plasma processes responsible for the microscopic structure of collisionless shocks. A tiny fraction of the particles crossing the shock is injected into the non-thermal energetic component that could get a substantial part of the ram pressure power dissipated at the shock. The energetic particles penetrate deep into the shock upstream producing an extended shock precursor. Scaling relations for postshock ion temperature and entropy as functions of shock velocity in strong collisionless multi-fluid shocks are discussed. We show that the multi-fluid nature of collisionless shocks results in excessive gas compression, energetic particle acceleration, precursor gas heating, magnetic field amplification and non-thermal emission. Multi-fluid shocks provide a reduced gas entropy production and could also modify the observable thermodynamic scaling relations for clusters of galaxies. ", "introduction": "\\label{Introduction} The observed large scale structure of the Universe is thought to be due to the gravitational growth of density fluctuations in the post-inflation era. In this model, the evolving cosmic web is governed by non-linear gravitational growth of the initially weak density fluctuations in the dark energy dominated cosmology. The web is traced by a tiny fraction of luminous baryonic matter. Cosmological shock waves are an essential and often the only way to power the luminous matter by converting a fraction of gravitational power to thermal and non-thermal emissions of baryonic/leptonic matter. At high redshifts ($z> 1100$) the pre-galactic medium was hot, relatively dense, ionised, with a substantial pressure of radiation. The cosmic microwave background (CMB) observations constrain the amplitudes of density inhomogeneities to be very small at the last scattering redshift $z \\sim$ 1000. Strong non-linear shocks are therefore unlikely at that stage. The universe expands, the matter cools, and eventually recombines, being mostly in neutral phase during the \"dark ages\" of the universe. At some redshift, $6 < z < 14$, hydrogen in the universe is reionised, likely due to UV radiation from the first luminous objects, leaving the intergalactic medium (IGM) highly reionised (see e.g. \\citet{Fan_ea06} for a recent review). The reionisation indicates the formation of the first luminous objects at the end of the \"dark ages\", either star-forming galaxies or Active Galactic Nuclei (AGN). The compact luminous objects with an enormous energy release would have launched strong (in some cases, relativistic) shock waves in the local vicinity of the energetic sources. At the same evolution stage, formation of strong density inhomogeneities in the cosmic structure occurs. Since then the non-linear dynamical flows in the vicinity of density inhomogeneities would have created large scale cosmic structure shocks of modest strength, thus heating the baryonic matter and simultaneously producing highly non-equilibrium energetic particle distributions, magnetic fields and electromagnetic emission. Most of the diffuse X-ray emitting matter was likely heated by cosmological shocks of different scales. Accretion and merging processes produce large-scale gas shocks. Simulations of structure formation in the Universe predict that in the present epoch about 40~\\% of the normal baryonic matter is in the Warm-Hot Intergalactic Medium (WHIM) at overdensities $\\delta \\sim 5-10$ (e.g. \\citealt{CenO99,Dave_ea01}). The WHIM is likely shock-heated to temperatures of 10$^5 -- 10^7$~K during the continuous non-linear structure evolution and star-formation processes. The statistics of cosmological shocks in the large-scale structure of the Universe were simulated in the context of the $\\Lambda$CDM-cosmology using PM / Eulerian adiabatic hydrodynamic codes (e.g. \\citealt{Miniati_ea00,Ryu_ea03,Kang_ea07}) and more recently with a smoothed particle hydrodynamic code by \\citet{Pfrommer_ea06}. They identified two main populations of cosmological shocks: $(i)$ high Mach number \"external\" shocks due to accretion of cold gas on gravitationally attracting nodes, and $(ii)$ moderate Mach number ($2 \\leq {\\cal M}_{\\rm s} \\leq 4$) \"internal\" shocks. The shocks are due to supersonic flows induced by relaxing dark matter substructures in relatively hot, already shocked, gas. The internal shocks were found by \\citet{Ryu_ea03} to be most important in energy dissipation providing intercluster medium (ICM) heating, and they were suggested by \\citet{Bykov_ea00cl} to be the likely sources of non-thermal emission in clusters of galaxies. Hydrodynamical codes deal with N-body CDM and single-fluid gas dynamics. However, if a strong accretion shock is multi-fluid, providing reduced post-shock ion temperature and entropy, then the internal shocks could have systematically higher Mach numbers. Space plasma shocks are expected to be {\\sl collisionless}. Cosmological shocks, being the main gas-heating agent, generate turbulent magnetic fields and accelerate energetic particles via collisionless multi-fluid plasma relaxation processes thus producing non-thermal components. The presence of these non-thermal components may affect the global dynamics of clusters of galaxies \\citet{Ostriker_ea05} and the $\\sigma_{\\rm v}$-$T$, $M$-$T$, $L_{\\rm X}$-$T$ scaling relations \\citet{Bykov05}. Detailed discussion of the cosmological simulations of the scaling relations with account of only thermal components can be found in \\citealt{13_borgani2008} - Chapter 13, this volume. In Sect.~\\ref{colls} we discuss the basic features of the standard collisional shocks. The main part of the review is devoted to physical properties of cosmological shocks with an accent on collisionless shocks and associated non-thermal components. In Sect.~\\ref{CRS} we discuss the most important features of multi-fluid collisionless shocks in the cosmological context including the effects of reduced entropy production, energetic particle acceleration and magnetic field amplification in the shocks. ", "conclusions": "Cosmological shocks convert a fraction of the energy of gravitationally accelerated flows to internal energy of the gas. They heat and compress the gas and can also accelerate energetic non-thermal particles and amplify magnetic fields. We discussed some specific features of cosmological shocks. $\\bullet$ The standard Rankine-Hugoniot relations based on the conservation laws for a steady single-fluid MHD shock allow to calculate the state of the fluid behind the shock once the upstream state and the shock strength are known. The coplanarity theorem for a plane ideal MHD shock states that the upstream and downstream bulk velocities, magnetic fields and the shock normal all lie in the same plane. $\\bullet$ Cosmological plasma shocks are likely to be collisionless as many other astrophysical shocks observed in the heliosphere and in supernova remnants. We review the basic plasma processes responsible for the microscopic structure of collisionless shocks. $\\bullet$ Collisionless shock heating of ions results in a non-equilibrium state just behind a very thin magnetic ramp region with a strongly anisotropic quasi-Maxwellian ion distributions. The possibility of collisionless heating of electrons by electromagnetic fluctuations in the magnetic ramp region depends on the extension of the fluctuation spectra to the electron gyro-scales, and could depend on the shock Mach number. Then the Coulomb equilibration processes are operating on the scales much larger than the collisionless shock width. $\\bullet$ Extended MHD shock waves propagating in turbulent media could accelerate energetic particles both by Fermi type acceleration in converging plasma flows and by DC electric field in quasi-perpendicular shocks. If the acceleration is efficient, then the strong shock could convert a substantial fraction (more than 10~\\%) of the power dissipated by the upstream bulk flow to energetic particles (cosmic rays). The compression ratio $r_{\\rm tot}$ at such a shock can be much higher, while the ion temperature behind the shock $\\propto r_{\\rm tot}^{-2}$ and the post-shock entropy are lower, than that in a standard single fluid shock. The shock structure consists of an extended precursor and a viscous velocity jump (subshock) indicated in Fig.~\\ref{sketch}. $\\bullet$ Strong collisionless plasma shocks with an efficient Fermi acceleration of energetic particles could generate strong MHD waves in the upstream and downstream regions and strongly amplify the upstream magnetic fields. A distinctive feature of the shock is a predicted possibility of gas pre-heating in the far upstream region due to MHD wave dissipation, that can produce an extended filament of temperature $\\gtrsim$ 0.1 keV. $\\bullet$ Shock waves both from the cosmic web formation processes and those due to cluster merging activity can play an important role in clusters of galaxies. Direct evidences for such shocks, as traced by radio relics and the temperature jumps in X-ray observations havebeen found only in a small number of clusters, and thus we need more observations." }, "0801/0801.0447_arXiv.txt": { "abstract": "We present equivalent width measurements and limits of six diffuse interstellar bands (DIBs, $\\lambda$4428, $\\lambda$5705, $\\lambda$5780, $\\lambda$5797, $\\lambda$6284, and $\\lambda$6613) in seven damped {\\Lya} absorbers (DLAs) over the redshift range $0.091 \\leq z \\leq 0.524$, sampling $20.3 \\leq \\log N({\\HI}) \\leq 21.7$. DIBs were detected in only one of the seven DLAs, that which has the highest reddening and metallicity. Based upon the Galactic DIB--$N({\\HI})$ relation, the $\\lambda$6284 DIB equivalent width upper limits in four of the seven DLAs are a factor of 4-10 times below the $\\lambda$6284 DIB equivalent widths observed in the Milky Way, but are not inconsistent with those present in the Magellanic Clouds. Assuming the Galactic DIB--$E(B-V)$ relation, we determine reddening upper limits for the DLAs in our sample. Based upon the $E(B-V)$ limits, the gas-to-dust ratios, $N({\\HI})/E(B-V)$, of the four aforementioned DLAs are at least $\\sim 5$ times higher than that of the Milky Way ISM. The ratios of two other DLAs are at least a factor of a few times higher. The best constraints on reddening derive from the upper limits for the $\\lambda$5780 and $\\lambda$6284 DIBs, which yield $E(B-V) \\leq 0.08$ for four of the seven DLAs. Our results suggest that, in DLAs, quantities related to dust, such as reddening and metallicity, appear to have a greater impact on DIB strengths than does {\\HI} gas abundance; the organic molecules likely responsible for DIBs in DLA selected sightlines are underabundant relative to sightlines in the Galaxy of similarly high $N({\\HI})$. With regards to the study of astrobiology, this could have implications for the abundance of organic molecules in redshifted galaxies. However, since DLAs are observed to have low reddening, selection bias likely plays a role in the apparent underabundance of DIBs in DLAs. ", "introduction": "Since their discovery by \\citet{hege22}, several hundred diffuse interstellar bands (DIBs) have been studied \\citep{gala00,jenn94,tuai00,wese00,hobbs07}, and yet no positive identifications of the carriers have been made. The DIBs span the visible spectrum between 4000 and 13,000~{\\AA}. Despite no positive identifications, several likely organic molecular candidates have emerged as the sources of the DIBs, including polycyclic aromatic hydrocarbons (PAHs), fullerenes, long carbon chains, and polycyclic aromatic nitrogen heterocycles (PANHs) \\citep[e.g.,][]{herb95, snow01, cox06a, hudg05}. The organic-molecular origin of the DIBs may give them an importance to astrobiology; PAHs are now considered an important early constituent to the inventory of organic compounds on Earth \\citep{bada02}. Via their infrared emitting vibrational bands, PAHs have been observed in high redshift dusty ultra-luminous infrared galaxies \\citep[ULIRGs, e.g.,][]{yan05}. Searching for DIBs using the technique of quasar absorption lines provides a different approach for charting the presence of possible organics to high redshift. As such, observing DIBs in high redshift galaxies may offer an independent method for constraining the environmental conditions in early-epoch galaxies governing the abundances of organic molecules, determining the cosmic epoch at which organic molecules first formed, and ultimately charting their evolution with redshift. Aside from the hundreds of detections within the Galaxy \\citep[e.g.][]{gala00,jenn94,tuai00,wese00,hobbs07}, DIBs have been detected in the Magellanic Clouds \\citep{welt06,cox06b,cox07}, M31 \\citep{cord08}, seven starburst galaxies \\citep{heck00}, active galaxy Centaurus A via supernova 1986A \\citep{rich87}, spiral galaxy NGC 1448 via supernovae 2001el and 2003hn \\citep{soll05}, one damped {\\Lya} absorber (DLA) at $z=0.524$ toward the quasar (QSO) AO~0235+164 \\citep{junk04, lawt06}, and one $z=0.157$ {\\CaII} selected absorber toward QSO J0013--0024 \\citep{elli07}. There are several environmental factors, such as {\\HI} column density \\citep{herb95,welt06}, reddening \\citep{welt06}, and metallicity and ionizing radiation \\citep{cox07}, that are related to DIB strengths. In the Galaxy, DIB absorption strengths correlate strongly with $N({\\HI})$ \\citep{herb95,welt06}. However, in the Magellanic Clouds, DIBs are weaker by factors of 7-9 (LMC) and $\\sim20$ (SMC) compared to those observed in the Galaxy with similar $N({\\HI})$ \\citep{welt06}. It is not known whether other galaxies in the Local Group and beyond obey the Galactic DIB--$N({\\HI})$ relation, or, like the Magellanic Clouds, show departures from this relation. Galaxies with high $N({\\HI})$ observed in absorption (i.e., DLAs) that reside at low to intermediate redshifts (where the prominent DIBs fall in the optical region) provide excellent astrophysical laboratories with which to investigate this issue. In this paper we search for $\\lambda$4428, $\\lambda$5780, $\\lambda$5797, $\\lambda$6284, and $\\lambda$6613 DIB absorption in seven low to intermediate redshift DLAs. In \\S~\\ref{sec:sample}, we give a brief summary of each intervening DLA in our sample. In \\S~\\ref{sec:obs}, we discuss the spectroscopic observations and data reduction of the background QSOs. In \\S~\\ref{sec:analysis}, we explain the procedure of our analysis, and the resulting spectra. In \\S~\\ref{sec:results}, we present our results and compare our data to the Galactic DIB--$N({\\HI})$ relation, deduce upper limits for the reddening, $E(B-V)$, determine lower limits on the gas-to-dust ratios, and discuss the role of metallicity for our sample of DLAs. We conclude in \\S~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} In this paper we employ a generalized method of the \\citet{schn93} technique to find lines and determine conservative equivalent width limits for DIBs in seven DLAs along with an assessment of uncertainties in these limits. We find: (1) The $\\lambda$6284 DIB in four of the DLAs in our sample have equivalent width upper limits that are 4--10 times lower then expected for similar $N({\\HI})$ relative to Galactic sight lines. These limits are not inconsistent with the $\\lambda$6284 DIB strengths found in the LMC and SMC. (2) Assuming the $\\lambda$5780 and $\\lambda$6284 DIB--$E(B-V)$ relations hold for DLAs, as it does for Galactic and Magellanic Cloud sightlines, we estimated upper limits on the reddening for our sample. In four of our DLAs we estimate $E(B-V) \\leq 0.08$ with two DLAs having an $E(B-V) \\leq 0.05$. These results are consistent with high redshift DLA samples, which have $E(B-V) < 0.04$ \\citep{elli05} and $E(B-V) < 0.02$ \\citep{murp04}. (3) Applying the $E(B-V)$ limits, one of our DLAs is consistent with having the same or larger gas-to-dust ratio as the SMC. Two of our DLAs are consistent with having gas-to-dust ratios at least as large as sightlines in the LMC. Three of our DLAs have less stringent limits that give lower limit gas-to-dust ratios consistent with Galactic, LMC, or SMC sightlines. The AO~0235+164 DLA has a measured $E(B-V)$ and $N({\\HI})$ that puts its gas-to-dust fraction on the high end of the LMC sightlines as stated in other work \\citep{junk04}. The $E(B-V)_{lim} \\leq 0.06$ constrained from the $\\lambda$6284 DIB is inconsistent with the known $E(B-V) = 0.23$ measured in \\citet{junk04}. We should have been able to detect the $\\lambda$6284 DIB in the AO~0235+164 DLA given the limit adopted by \\citet{lawt06}. Three possibilities are that our method of determining the $E(B-V)_{\\rm lim}$ does not apply to DLAs, our limits do not adequately take into account the large atmospheric absorption band (see Fig.~\\ref{s:0235}$b$), or that conditions are not favorable to the formation or survival of this DIB carrier. It is interesting to speculate whether ionization conditions may be an important factor in inhibiting or enhancing DIB strengths. \\citet{welt06} test the ionization effects of DIBs along Galactic, LMC, and SMC lines of sight. However, they do not find any significant trends. On the contrary, \\citet{cox07} measure the UV radiation field along lines of sight toward the SMC and find UV radiation is an important environmental factor in DIB strengths. Laboratory spectroscopists claim that the DIBs may be due to partially ionized PAHs because they produce a wealth of absorption features in the optical spectrum \\citep{snow01}. If this is the case then a significant UV radiation may be required. However, UV radiation that is too high will dissociate the molecules. In this work, we are unable to explore the affects of ionization conditions; knowledge of ionization conditions in specific DLAs is difficult to obtain. It is also interesting to speculate whether metallicity may be an important factor in inhibiting or enhancing DIB strengths. However, since we have robust metallicity measurements for only two of the DLAs in our sample, we cannot directly address the affects of metallicity. We do point out that obtaining deep spectra of DIBs in DLAs with known metallicity could be a fruitful future research direction, especially if reddening were also known. A direct comparison between the low metallicity Q1229--020 DLA and the high metallicity AO~0235+164 DLA might be fruitful as a first examination of the affects of metallicity in determining what governs DIB strengths in DLAs. \\subsection{Implications of our Results} Our results imply that reddening is a more crucial indicator of DIB strengths than is {\\HI} content for DLAs; the low reddening in DLA selected galaxies inhibits the presence of DIBs. However, our sample is small and additional observations are required to ascertain the strength of this statement (metallicity and ionization conditions likely play a role as well). The weakness of DIB strengths in our sample hints that the environments of high $N({\\HI})$ DLA-selected galaxies may be less suitable to create and/or sustain the organic molecules than those of the Galaxy. Not only is the immediate solar environment beneficial for sustaining life, but it may be that the Galaxy is a more hospitable location for the survival of organic molecules that may have been the precursors to biology on Earth. If these molecules are important as precursors to life in the universe, charting their presence to high redshift places constraints on how long ago and in which environments life could have potentially formed. The presence of DIBs in the AO~0235+164 DLA demonstrates that organic molecules existed in at least one DLA selected environment at a redshift of $z \\sim 0.5$, or some 5 Gyrs ago \\citep{lawt06}. Due to their weakness relative to DIBs in the Galaxy, observing DIBs in the general population of high redshift galaxies (as opposed to ULIRGs and star bursting galaxies) remains a challenge. Selecting galaxies by DLA absorption may yet prove to be a lucrative method for detecting DIBs at high redshifts, provided the detection sensitivity can be increased. However, most known DLAs reside at high redshift where the DIBs move into the near-IR, where high sensitivity spectroscopy is time intensive." }, "0801/0801.2504_arXiv.txt": { "abstract": "In the framework of nonextensive statistical mechanics, the equilibrium structures of astrophysical self-gravitating systems are stellar polytropes, parameterized by the polytropic index $n$. By careful comparison to the structures of simulated dark-matter halos we find that the density profiles, as well as other fundamental properties, of stellar polytropes are inconsistent with simulations \\textit{for any value of $n$}. This result suggests the need to reconsider the applicability of nonextensive statistical mechanics (in its simplest form) to equilibrium self-gravitating systems. ", "introduction": "Nonextensive statistical mechanics is a generalization of thermodynamics and statistical mechanics proposed by C.\\ Tsallis in 1988 \\cite{tsallis88}, aiming to overcome the limitations of the Boltzmann entropy in its conventional applications \\cite{karlin02,taruya03b,chavanis05,gross05}. The theory has considerably widened the fields of application of statistical physics, allowing the description of systems affected by nonlocal effects, such as long-range forces and memory effects. The study of astrophysical self-gravitating systems was one of the first applications of the theory \\cite{plastino93}. This approach has recently witnessed a renewed interest in the hope that it may provide a theoretical basis for the description of the universal structure of dark-matter (DM) halos \\cite{hansen05,leubner05,kronberger06,zavala06}. Nonextensive statistical mechanics predicts their equilibrium states to be stellar polytropes (SP) \\cite{plastino93,taruya03}, which have been claimed to fit DM halos as well as the usual Navarro, Frenk \\& White model \\cite{kronberger06,zavala06} (hereafter NFW \\cite{NFW96}). Moreover, SP have the distinct advantage of being analytically derived from nonextensive statistical mechanics, while most models describing DM halos are empirical fits to N-body simulations. In this context, nonextensive statistical mechanics appears as an attractive framework for providing a theoretical understanding of the structure of DM halos and self-gravitating systems in general. In this paper we compare, in a parameter independent way, the equilibrium configuration of astrophysical self-gravitating systems predicted by nonextensive statistical mechanics to simulated DM halos. We clarify the issue of which central boundary conditions to use for comparing SP to simulated halos. On this basis, we establish that simulated DM halos do not corroborate the predictions of nonextensive statistical mechanics. This result calls into question the direct applicability of nonextensive statistical mechanics to equilibrium astrophysical self-gravitating systems. ", "conclusions": "We have established here that the predictions of nonextensive statistical mechanics are not corroborated by simulations of DM halos. Our results are based on readily observable quantities, i.e., the matter density itself, and profiles derived from it. These findings are in direct contrast to previous works which found reasonable agreement because they either considered only the outer parts of simulated halos \\cite{leubner05}, or used a non-zero density slope as initial condition near the center \\cite{kronberger06}, or used values of $n<5$ that lead to too steep density profiles in the outer parts \\cite{zavala06}. Support for our conclusions, however, comes from Barnes et al.\\ \\cite{barnes07} who, even though fixing a non-zero density slope near the center to match the NFW density profile, found inconsistency between SP and DM halo velocity dispersion profiles. While we are cautious about the physical relevance of such profiles, it is of interest to emphasize that nor cored neither cuspy polytropes can describe properly simulated DM halos. These results imply a fundamental difference in the matter distribution of astrophysical self-gravitating systems predicted by nonextensive statistical mechanics and by N-body simulations. Such a discrepancy raises three main questions. Is our comparison with simulated DM halos valid? Is nonextensive statistical mechanics the proper theory to describe collisionless long-range interaction systems at equilibrium? And how does this study relate to observed DM halos? We briefly address these issues below. 1a. We study idealized systems which do not take into account complex effects in DM halos such as velocity anisotropy \\cite{hansen06c} or triaxiality \\cite{Hayashi07}. However, the error we make by using isotropic and spherically averaged models to represent DM halos is small compared to the discrepancy we find between SP and simulated halos. Moreover, we do not address here general problems of statistical mechanics of self-gravitating systems like, e.g., infinite mass \\cite{hjorth91}, so as to focus on the issues specific to nonextensive statistical mechanics. 1b. N-body simulations depend on various parameters, such as the choice of the softening length and the number of particles, which can introduce numerical effects in the resulting halos. The softening of the gravitational force on small scale, introduced to avoid two-body interaction, creates a core at the center of the halo, approximately the size of the softening length. The number of particles fixes the maximum phase-space density resolved at redshift zero, and if not large enough, a core appears due to the lack of particles in the center. Both effects lead to a shallower density profile in the center of halos if the simulation is of low resolution \\cite{moore98}. Therefore, the discrepancy we observe between simulated profiles and SP is genuine and increases with the resolution of numerical simulations. 1c. We compare theoretical statistical predictions to cosmological simulations, but the latter depend on initial conditions and cosmological parameters. Initial conditions are fixed by the shape of the power spectrum of initial fluctuations $P(k) \\propto k^{n_s}$, i.e., by the choice of the index $n_s$. From CMB observations, we have an indication that $n_s=0.958 \\pm 0.016$ \\cite{spergel07}, but tests on cosmological simulations proved that they depend very little on the choice of $n_s$ and on the cosmological model used \\cite{navarro97}. Therefore, the universal profiles of simulated halos can be compared to statistical theories as a general prediction. 1d. The choice of a scaling is necessary to compare DM halos models, but can bring a visualization bias. We checked if the use of $r_{-2}$ leads us to overestimate the disagreement between nonextensive statistical mechanics and N-body simulations. However, adjusting the fit in the outer parts of the halo, e.g., scaling to the virial radius, increases even more the discrepancy in the inner parts, while scaling profiles to fit well at smaller radii makes the discrepancy appear in the outer parts of the halo. 2. Nonextensive statistical mechanics relies on three assumptions: \\textit{a)} the generalized entropy $S_q$ is the right form to describe long-range interaction systems; \\textit{b)} the system is at equilibrium, and its most probable state is given by the maximum entropy principle; \\textit{c)} $S_q$ is maximized at fixed energy, leading to a system whose distribution function depends only on the energy per unit mass, and has isotropic velocity dispersion. Though simulated halos show evidence of velocity anisotropy, the relation between density slope and velocity anisotropy \\cite{hansen06c} shows that it is isotropic in the center, where violent relaxation \\cite{lyndenbell67} (the process by which strong potential fluctuations efficiently drives the system towards equilibrium) is most effective. Therefore, the assumptions \\textit{b)} and \\textit{c)} hold true in the center. However, it is in the inner parts that the disagreement between SP and simulated halos is the strongest. Hence, we suggest that the assumption \\textit{a)}, i.e., the choice of the generalized entropy $S_q$ motivated by nonextensive statistical mechanics, cannot be used to predict the equilibrium structures of simulated DM halos. 3. Simulated DM halos are in good agreement with observations \\cite{pointecouteau05}, except for the inner core found from spiral galaxy rotation curves \\cite{gentile04}. Stellar polytropes, interestingly, have an inner core too, but the discrepancy we observe between SP and simulated DM halos is too large to explain the core of observed DM halos. Adding adiabatic contraction in simulations results in DM halos with steeper inner parts, which would not change our conclusions. In summary, we have established that nonextensive statistical mechanics \\cite{tsallis88, plastino93, taruya03}, a theory generalizing classical statistical mechanics and thermodynamics, does not describe the equilibrium state of astrophysical self-gravitating systems, as represented by DM halos formed in N-body simulations." }, "0801/0801.1111_arXiv.txt": { "abstract": "As first Paper of a series devoted to study the old stellar population in clusters and fields in the Small Magellanic Cloud, we present deep observations of NGC\\,121 in the F555W and F814W filters, obtained with the Advanced Camera for Surveys on the {\\it Hubble Space Telescope}. The resulting color-magnitude diagram reaches $\\sim 3.5$ mag below the main-sequence turn-off; deeper than any previous data. We derive the age of NGC\\,121 using both absolute and relative age-dating methods. Fitting isochrones in the ACS photometric system to the observed ridge line of NGC\\,121, gives ages of $11.8 \\pm 0.5$~Gyr (Teramo), $11.2 \\pm 0.5$~Gyr (Padova) and $10.5 \\pm 0.5$~Gyr (Dartmouth). The cluster ridge line is best approximated by the $\\alpha$-enhanced Dartmouth isochrones. Placing our relative ages on an absolute age scale, we find ages of $10.9 \\pm 0.5$~Gyr (from the magnitude difference between the main-sequence turn-off and the horizontal branch) and $11.5 \\pm 0.5$~Gyr (from the absolute magnitude of the horizontal branch), respectively. These five different age determinations are all lower by 2 -- 3 Gyr than the ages of the oldest Galactic globular clusters of comparable metallicity. Therefore we confirm the earlier finding that the oldest globular cluster in the Small Magellanic Cloud, NGC\\,121, is a few Gyr younger than its oldest counterparts in the Milky Way and in other nearby dwarf galaxies such as the Large Magellanic Cloud, Fornax, and Sagittarius. If it were accreted into the Galactic halo, NGC\\,121 would resemble the ``young halo globulars'', although it is not as young as the youngest globular clusters associated with the Sagittarius dwarf. The young age of NGC\\,121 could result from delayed cluster formation in the Small Magellanic Cloud or result from the random survival of only one example of an initially small number star clusters. ", "introduction": "Characterizing old stellar populations provide important constraints on the early star formation histories of galaxies. Only the satellite galaxies of the Milky Way (MW) are sufficiently close to resolve individual stars well below the oldest main-sequence turn-offs, which is a pre-condition for accurate photometric age dating of old stellar populations. All Local Group galaxies, for which adequate data exist, appear to contain stars older than 10~Gyr \\citep{grebel04}. This result is based on main-sequence turn-off photometry of globular clusters and field populations in Galactic satellites and a few more distant Local Group galaxies \\citep[e.g., ][]{brown07, cole07}, as well as the detection of horizontal branch stars (including RR Lyrae variables) in the Local Group and beyond \\citep[e.g., ][] {held00,harb01, saraj02, clementini03, pritzl04}. Globular clusters are preferred as the basis for old stellar population age tracers since they are usually single-age, single-metallicity objects facilitating comparative studies. Moreover, while globular cluster systems exhibit a range of ages \\citep[e.g., ][]{deang05}, the oldest ones may belong to the most ancient surviving stellar systems to have completed their formation in the youthful Universe \\citep[e.g., ][]{moore06}. In those nearby galaxies where relative age dating based on main-sequence photometry was carried out in comparison to the oldest globular clusters in the Milky Way, no age difference within the measurement accuracy was found \\citep[e.g., ][ and references therein]{grebel04, brown07, cole07}. The relative age dating of the oldest identifiable Population~II objects thus indicates a common epoch of substantial early star formation in the Milky Way and its companions, although information about a putative, even older Population III remains to be uncovered in these objects. A galaxy that may {\\em not}\\ share this common epoch of early star formation -- at least not with respect to its globular clusters \\citep[e.g., ][]{saraj98} -- is the Small Magellanic Cloud (SMC)\\footnote{There may be additional exceptions in more distant dwarf irregular galaxies regarding the common epoch of earliest Population~II star formation, although also these galaxies evidently contain old populations \\citep[e.g., ][]{grebel01, maka02}}. The SMC is one of the closest and therefore best studied dwarf galaxies orbiting our Galaxy. While the SMC hosts a large number of intermediate-age and young star clusters, it only contains one ''old'' globular cluster, NGC\\,121, which is also the most massive star cluster. NGC\\,121 is located $\\sim$ 2.4\\degr ($\\sim$3~kpc) west of the SMC bar at ($\\alpha_{J2000.0}$, $\\delta_{J2000.0}$) = ($0^h26^m47.0^s$, $-71\\arcdeg32'12.0''$). NGC\\,121 is the only cluster in the SMC that is sufficiently old to have developed an extended red horizontal branch \\citep{stry85} and to contain RR Lyrae stars. Indeed, whether or not to call a star cluster a globular cluster is a matter of definition. In this case we refer to \\citet{salgir02} who consider Lindsay\\,1 as having a stumpy red clump and not a red horizontal branch. Three RR Lyrae stars were discovered in NGC\\,121 by \\citet{thack58}. \\citet{grah75} found a fourth RR Lyrae variable in the cluster and an additional 75 in a 1 $\\times$ 1.3 square degree field centered on NGC\\,121. Studies of various clusters in the LMC and in the MW showed that the presence of RR Lyrae variables indicates that the parent population is as old as or older than $\\sim$ 10~Gyr. An important question is whether NGC\\,121 is as old as the typical old globular clusters in the Large Magellanic Cloud (LMC) and in the MW. Previous studies found ages ranging from 8 to 14~Gyr for NGC\\,121 \\citep{stry85, Walker91, migh98, udal98, shara98, dol01} using a variety of different techniques. Studies based on the deepest available color-magnitude diagrams from Hubble Space Telescope (HST) observations with the Wide Field and Planetary Camera 2 (WFPC2) indicate an age of 10 to 10.6~Gyr for NGC\\,121, suggesting that this globular cluster is several Gyr younger than the oldest globulars in other nearby galaxies and in the MW \\citep{shara98, dol01}. The capabilities of the Advanced Camera for Surveys (ACS) provide an improvement in both sensitivity (depth) as well as angular resolution, which is essential for a reliable photometric age determination in this dense star cluster. Here we present deep photometry of NGC\\,121 obtained with ACS aboard the HST. We determine the age of NGC\\,121 utilizing both absolute and relative methods \\citep[e.g., ][]{chab96Sci}. The current study is the first in a series of papers based on HST studies of rich intermediate-age and old star clusters in the SMC. In addition to NGC\\,121, six intermediate-age SMC star clusters have been observed as part of our program: Lindsay\\,1, Kron\\,3, NGC\\,339, NGC\\,416, Lindsay\\,38 and NGC\\,419. We will derive fiducial ridgelines and fit isochrones to obtain accurate ages for each cluster using the same reduction techniques and isochrone models as described here (see $\\S$~\\ref{sec:obs}-\\ref{sec:age}), and will present our results in future papers. In Table~\\ref{tab:journalobs} we list the cluster identification, date of observation, passband, exposure times and location of all clusters in our HST program (GO-10396; principal investigator: J.~S.~Gallagher). In the next Section we describe the data reduction procedure. In $\\S$~\\ref{sec:CMD} we present the color-magnitude diagram (CMD) of NGC\\,121 and discuss its main features. In $\\S$~\\ref{sec:age} we describe our age derivation methods and present our results. ", "conclusions": "We derived ages for the old SMC globular cluster NGC\\,121 based on our high dynamic range HST/ACS photometry that extends at least three magnitudes below its MSTO. In order to obtain absolute ages, we applied three different isochrone models. These isochrone models yielded ages of $11.2 \\pm 0.7$~Gyr (Padova), $11.8 \\pm 0.5$~Gyr (Teramo), and $10.5 \\pm 0.5$~Gyr (Dartmouth). We find the $\\alpha$-enhanced Dartmouth isochrones provide the closest approximation to the MS, SGB, and RGB, whereas the other models cannot reproduce the slope of the upper RGB. High-resolution spectroscopy indicates that NGC\\,121 is indeed $\\alpha$-enhanced \\citep{john04}, a property that it shares with many of the old outer Galactic halo globulars. Given the proximity of NGC\\,121 to the SMC on the sky and its distance, its physical association with the SMC seems well-established. Our determinations of relative ages for NGC\\,121 are consistent with the results of our absolute age determination. Relative age estimates, when converted to an absolute age scale, are $10.9 \\pm 0.5$~Gyr ($\\Delta V_{TO}^{HB}$), $10.8 \\pm 1.0$ ($M_V(HB)$) and $11.5 \\pm 0.5$~Gyr ($M_{V(BTO)}$). These numbers agree well with the absolute age derivations. Our results confirm that NGC\\,121 is 2--3~Gyr younger than the oldest MW and LMC clusters (as also found in earlier WFPC2 studies). NGC\\,121 is similar in age to the youngest globular cluster in the Fornax dSph \\citep{buon99}, and to several of the young Galactic halo clusters. On the other hand, NGC\\,121 is not as young as some of the Sgr dwarf galaxy's globular clusters or the youngest Galactic globular clusters. It is intriguing that the SMC -- in contrast to other Galactic companion dwarf galaxies with globulars -- does not contain any old classical globular clusters. But given the existence of only one cluster and the question of star cluster survival, this could be a result of the one survivor from the SMC's epoch of globular cluster formation randomly sampling an initial distribution of star cluster ages. On the other hand, in low-mass galaxies without bulges, spiral density waves, and shear it is much more difficult to destroy globular clusters through external effects. That this cluster is both younger than the Galactic mean and enhanced in $\\alpha$-elements may have interesting implications for the early development of the SMC. It also is intriguing that the only globular cluster in the SMC is not very metal-poor. The SMC must have experienced substantial enrichment prior to the formation of NGC\\,121. In the LMC, where two main epochs of the formation of populous compact star clusters have been found \\citep[e.g., ][]{Bertelli92}, a few globular clusters are found that are old enough to exhibit blue HBs. Interestingly, these globular clusters, which are similarly old as the oldest Galactic globulars \\citep{olsen98}, have a similar metallicity to NGC 121 \\citep{john04} (e.g., NGC\\,1898, NGC\\,2019), indicating very early chemical enrichment. The MW also contains old classical globular clusters (with blue HBs) that have similarly high metallicities as the somewhat younger NGC 121. Evidently, the conditions for and the efficiency of star formation varied in these three galaxies at early epochs. After NGC 121 formed there was a hiatus in surviving stars clusters and thus possibly in cluster formation activity in the SMC: The second oldest SMC cluster is Lindsay 1 with an age of $\\sim$ 8~Gyr (Glatt et al. 2007, in preparation). Since then compact populous star clusters formed fairly continuously until the present day in the SMC (e.g., Da Costa 2002) -- in contrast to both the LMC and the MW. In forthcoming papers on our ACS photometry of SMC clusters and field populations we will explore the evolutionary history of the SMC in more detail. Clearly, clues about the early star formation history of the SMC will have to come from its old field populations." }, "0801/0801.4734_arXiv.txt": { "abstract": "The observed velocities of pulsars suggest the possibility that sterile neutrinos with mass of several keV are emitted from a cooling neutron star. The same sterile neutrinos could constitute all or part of cosmological dark matter. The neutrino-driven kicks can exhibit delays depending on the mass and the mixing angle, which can be compared with the pulsar data. We discuss the allowed ranges of sterile neutrino parameters, consistent with the latest cosmological and X-ray bounds, which can explain the pulsar kicks for different delay times. ", "introduction": " ", "conclusions": "" }, "0801/0801.3863_arXiv.txt": { "abstract": "Supergiant Fast X-ray Transients are obviously related to persistent Supergiant X-ray Binaries. Any convincing explanation for their behaviour must consistently take into account all types of X-ray sources powered by wind accretion. Here we present a common framework for wind accreting sources, within the context of clumpy wind models, that allows a coherent interpretation of their different behaviours as an immediate consequence of diverse orbital geometries. ", "introduction": "Traditionally, High Mass X-ray binaries (HMXBs) have been divided in three classes: ``classical'' bright ($L_{{\\rm X}}\\sim10^{38}\\:{\\rm erg}\\,{\\rm s}^{-1}$) sources fed by incipient localised Roche lobe overflow, Be/X-ray binaries, presenting different behaviours but always associated to the presence of a circumstellar decretion disk, and wind-fed Supergiant X-ray Binaries (SGXBs). In the last group, the compact object is fed by accretion from the strong radiative wind of an OB supergiant, leading to a persistent X-ray source with $L_{{\\rm X}}\\sim10^{36}\\:{\\rm erg}\\,{\\rm s}^{-1}$, displaying large variations on short timescales, but a rather stable flux in the long run. The compact object is a relatively close ($P_{{\\rm orb}}\\la15\\:{{\\rm d}}$) orbit with low or negligible eccentricity. Over the last few years, there has been a huge increase in the number of HMXBs known \\citep[e.g.,][]{wal06}. Many new sources have been found to display very brief outbursts, with a rise timescale of tens of minutes and lasting only a few hours \\citep{sgue05}. Several have been identified with OB supergiants: \\citep[e.g., the prototype XTE~J1739$-$302;][]{smith06,neg06b}. The distances to their counterparts imply typical $L_{{\\rm X}}\\sim10^{36}\\:{\\rm erg}\\,{\\rm s}^{-1}$ at the peak of the outbursts. A number of other systems have been shown to have similar X-ray behaviours leading to the definition of a class of Supergiant Fast X-ray Transients (SFXTs) \\citep{esa,smith06}. Around 12 such systems are known \\citep{wzh07}. It is perhaps possible to divide them in two groups, the first one characterised by very low quiescence $L_{{\\rm X}}$ and high variability and the second one with higher average $L_{{\\rm X}}$ and smaller variability factors. The second group, then, could be thought of as persistent SGXBs with an average $L_{{\\rm X}}$ below the canonical $\\sim10^{36}\\:{\\rm erg}\\,{\\rm s}^{-1}$ over which flares are superimposed \\citep{wzh07}. To some degree, this distinction may be due to observational biases (cf. the case of XTE~J1739$-$302; Blay et al., submitted). In any case, it is clear that the separation between SFXTs and SGXBs is not well defined. Two other systems have shown flaring only during episodes separated by regular intervals: IGR J00370+6122 \\citep{zand07} and IGR J11215-5952 \\citep{sidoli}. This modulation is most likely related to an orbital period. Because of their very different properties, we do not count these two systems as SFXTs, but they are also wind-fed accretors and any theory attempting an explanation to the behaviour of SFXTs must take into account the existence of other wind accretors. At least two SFXTs have been observed to increase their $L_{{\\rm X}}$ by a factor $>100$ (going from deep quiescence to outburst) in only a few minutes: XTE~J1739$-$302 \\citep{sak02} and IGR~J17544$-$2619 \\citep{zand05}. AX~J1841.0$-$0535 showed a $>10$ increase in count-rate over $\\la 1\\:$h \\citep{bam01}. Such sharp rises are incompatible with any explanation of their behaviour based solely on orbital motion through a smooth medium. ", "conclusions": "" }, "0801/0801.3264_arXiv.txt": { "abstract": "We present an abundance analysis based on high dispersion and high signal-to-noise ratio Keck spectra of a very highly microlensed Galactic bulge dwarf, \\mystar, with \\teff $\\sim 5400$~K. The amplification at the time the spectra were taken ranged from 350 to 450. This bulge star is highly enhanced in metallicity with [Fe/H]\\footnote{We adopt the usual spectroscopic notations that [A/B] ~ $\\equiv ~ log_{10} (N_A/N_B)_* - log_{10} (N_A/N_B)_{\\odot}$, and that log$[\\epsilon(A)] ~ \\equiv ~ log_{10} (N_A/N_H) + 12.00$, for elements $A$ and $B$.} = \\fehogle ~dex. The abundance ratios for the 28 species of 26 elements for which features could be detected in the spectra are almost all solar. In particular, there is no evidence for enhancement of any of the $\\alpha$-elements including O and Mg. We conclude that the high [Fe/H] seen in this star, when combined with the equally high [Fe/H] derived in previous detailed abundance analysis of two other Galactic bulge dwarfs, both also highly magnified by microlensing, implies that the median metallicity in the Galactic bulge is very high. We thus infer that many previous estimates of the metallicity distribution in the Galactic bulge have substantially underestimated the mean Fe-metallicity there due to sample bias, and suggest a candidate mechanism for such. If our conjecture proves valid, it may be necessary to update the calibrations for the algorithms used by many groups to interpret spectra and broad band photometry of the integrated light of very metal-rich old stellar populations, including luminous elliptical galaxies. ", "introduction": "Microlensing occurs when a ``lens'' (star, planet, black hole, etc) becomes closely aligned with a more distant ``source'' star, whose image it both magnifies and distorts. Normally the observer is most interested to learn about the lens, but microlensing data can simultaenously serve as a powerful probe of the source. If the source transits the lens (or the ``caustics'' generated by the lens, in the case of binary lenses), then it resolves the source, allowing detailed limb-darkening profiles from a photometric times series \\citep{fields03} or even spectral resolution of surface features like the chromosphere \\citep{cassan04} from judiciously taken spectra. The lens can also serve simply to amplify the role of the telescope as a ``light bucket''. \\citet{minniti98} provocatively titled their report on observations of a bulge dwarf that was magnified by a factor $A=2.25$ ``Using Keck I as a 15m Diameter Telescope'' and \\cite{cavallo} presented a preliminary abundance analysis for six stars with small magnification ($2.5 < A < 30$), including this star and two other bulge dwarfs. Obviously, this trick could in principle be improved to arbitrary ``diameters'' simply by observing the events at higher magnification. The problem is that it is extremely difficult to recognize high-magnification events in advance, and harder still to activate large-telescope observations in time to take advantage of them. The process has been facilitated by microlensing planet hunters, who prize high-magnification events because of their extreme sensitivity to planets \\citep{ob05071,ob05169,ob06109}. \\citet{johnson07} were the first to piggy-back on the microlensing planet hunters, obtaining a 15 minute Keck spectrum of the Galactic bulge dwarf \\johnstar\\ at magnification $A=135$. This star, the first bulge dwarf with a high-quality spectrum, proved to be extremely metal-rich, a fact that might be telling us that the bulge is much more metal-rich than seems indicated by available spectra of giants, but could also just be the ``luck of the draw''. The \\citet{johnson07} results therefore substantially raised the premium on obtaining highly-magnified spectra of bulge dwarfs. On 2 July 2007, the OGLE collaboration\\footnote{http://www.astrouw.edu.pl/$\\sim$ogle/ogle3/ews/ews.html} \\citep{ews} announced OGLE--2007--BLG--349\\ (RA=18:05:24.43; DEC = $-26$:25:19.0) at Galactic coordinates $(l,b)=(4.4,-2.5)$, i.e., 5.1$^{\\circ}$ from the Galactic center, as a probable microlensing event. The Microlensing Follow Up Network\\footnote{http://www.astronomy.ohio-state.edu/$\\sim$microfun/} ($\\mu$FUN) began monitoring the event on 18 August to determine whether the event would be high-magnification and on 3 September issued a general alert that it would reach at least $A>200$ two nights hence. On this basis, $\\mu$FUN organized world-wide photometric observations, whose outcome will be reported elsewhere \\citep{dong08} and also contacted JGC at the Keck telescope to recommend intensive observations. The ability to obtain high resolution, high quality spectra of Galactic bulge stars and to carry out a detailed abundance analysis offers an unbiased way to determine the metallicity distribution of stars in the Galactic bulge, as well as their detailed chemical inventory. Our abundance analysis of \\mystar\\ is described in the first few sections of the paper, with the key results presented in \\S\\ref{section_abund_results}. In an effort to explain our rather surprising results, we indicate in \\S\\ref{section_discussion} how past studies of the brightest giants, which are the only bulge stars for which detailed abundance analyses can be derived from spectra obtained under normal conditions, may be subject to previously ignored selection effects, and how this might impact studies of the integrated light for metal-rich old simple stellar populations. Abundance ratios in the Galactic bulge giants and in the microlensed dwarfs are discussed in \\S\\ref{section_abund_ratios}. A brief summary concludes the paper, while an appendix discusses the behavior of selected diffuse interstellar bands in the spectrum of \\mystar. ", "conclusions": "} An early attempt to determine the metallicity distribution in the Galactic bulge was that of \\cite{sadler96}, who used low resolution spectroscopy for 268 bulge giants and red clump stars to derive a mean [Fe/H] of $-0.11\\pm0.04$~dex. \\cite{ramirez00} studied a sample of M giants in the near-IR; they found a very similar mean [Fe/H] of $-0.21$~dex with a dispersion of 0.30~dex. \\cite{zocalli} used extensive optical and near-IR photometry with CMD fitting; they derived a somewhat lower mean [Fe/H]. In all these cases, the calibration of the metallicity scale relied on Galactic globular clusters. \\cite{zocalli} suggest that the differences between several of these studies depend crucially on the abundances adopted for the two highest metallicity GCs with high resolution abundance analyses, NGC~6553 and NGC~6558. [Fe/H] values for these two GCs have ranged over more than 0.5~dex in the literature, but since the work of \\cite{cohen99} and \\cite{carretta01}, who suggested values higher than most previous studies, more recent analyses have settled toward the higher values, see, e.g., \\cite{zoccali04} and \\cite{vlt_6553}. An early high dispersion spectroscopic study of Galactic bulge K giants is that of \\cite{mcwilliam94}, who found a mean [Fe/H] in their sample of $-0.25$~dex. \\cite{fulbright06} updates and expands upon this earlier work. They then use their detailed abundance analyses of 27 K giants in Baade's window to recalibrate the metallicities for the much larger samples of \\cite{sadler96} and of \\cite{rich88}. The mean [Fe/H] they thus deduce is $-0.10\\pm0.04$~dex. The median [Fe/H] of their sample is also sub-solar. As infrared echelle spectrographs have become available on 8-m class telescopes, high dispersion studies of Galactic bulge giants in the near-IR have become possible, see, e.g., \\cite{rich05}, \\cite{cunha06} and \\cite{rich07}. The mean [Fe/H] from the sample of M giants studied by \\cite{rich07} in the Galactic bulge is well below solar, with [Fe/H] $-0.22\\pm0.01$~dex. The median Fe-metallicity of the small sample of giants with near-IR spectra analyzed by \\cite{cunha06} is also slightly below solar metallicity. Since this sample largely overlaps that of \\cite{fulbright06} it is not included in the figures. Very recently, \\cite{zoccali08} have presented initial results of a survey of Fe-metallicity in the Galactic bulge from spectra with $\\lambda/\\Delta\\lambda = 20,000$ of about 800 stars. They find a radial gradient in [Fe/H] within the bulge with the mean value going from +0.03~dex at $b = -4^{\\circ}$ to $-0.12$~dex at $b = -6^{\\circ}$, and a sharp cutoff towards higher metallicities. All of these samples of Galactic bulge stars are of luminous giants and/or of red clump stars. They all have mean and median [Fe/H] values that are slightly sub-solar and are similar to the mean [Fe/H] of $-0.1$~dex found for local disk stars by \\cite{localdisk}. Yet \\mystar, analyzed here, and \\johnstar\\ \\citep{johnson07}, another Galactic bulge microlensed dwarf with a high quality detailed abundance analysis, both have [Fe/H] $\\sim +0.5$~dex; the third such star found to date, \\moa\\, with a lower signal-to-noise ratio spectrum, analyzed by \\cite{johnson08}, has a somewhat lower Fe-metallicity, [Fe/H] = $+0.36\\pm0.18$~dex. The comparison between the [Fe/H] distribution for the recalibrated sample of \\cite{sadler96} by \\cite{fulbright06} and the [Fe/H] values deduced for these three Galactic bulge dwarfs is shown in Figure~\\ref{figure_feh_hist}. While three stars is an uncomfortably small sample, it is difficult to believe that this is consistent with the published metallicity distributions for the Galactic bulge giants and/or red clump stars. The probability that the sample of three stars would have such high [Fe/H] values by chance is less than 1\\% given the [Fe/H] distributions of the larger of the relevant studies in the Galactic bulge claiming to have unbiased samples, including \\cite{ramirez00} with 110 M giants, which must be considered unbiased at the high metallicity end, \\cite{sadler96} (268 bulge stars), and the recalibrated verion of the latter by \\cite{fulbright06}. A KS test shows that the probability of finding the three carefully studied microlensed bulge dwarfs at their very high [Fe/H] values if the underlying metallicity distribution is that found by \\cite{zoccali08} from high dispersion spectra of a very large sample of giants in three bulge fields is less than 10$^{-4}$. This calculation takes into account the radial gradient in the metallicity distribution function found by \\cite{zoccali08}. We next assemble the evidence that \\mystar\\ is actually a dwarf in the Galactic bulge. We have already demonstrated that the microlensed bulge dwarf studied here has stellar parameters that, when combined with the unlensed magnitude from the mirolensing light curve, yield an unlensed luminosity consistent with that expected for a dwarf of its \\teff\\ at the distance of the Galactic center. This is true to within 0.2~mag for the other two such stars as well \\johnstar\\ and \\moa\\ \\citep{johnson07,johnson08}. The three microlensed bulge dwarfs studied to date come from a kinematically hot population; they have radial velocities of +99, $-$154 and +113 \\kms\\ with a typical uncertainty of less than 2~\\kms, thus showing a dispersion in $v_r$ consistent with that inferred from large samples of bulge giants \\citep[see, e.g.][]{sadler96,zoccali08}. About 4\\% of the dwarfs in the solar neighborhood from the proper motion sample of \\cite{grenon89} appear to be old and very metal-rich, with [M/H] $> +0.30$~dex. These stars are almost all on eccentric orbits with small pericenters ($\\lesssim 3$~kpc). \\cite{castro97} and \\cite{pompeia03} have carried out detailed abundance analyses of some of these stars to find that the most metal rich of them reach [Fe/H] +0.55~dex, and have solar abundance ratios in general. These local dwarfs are thus similar in their abundances to \\mystar. The references cited suggest that they are possibly stars on chaotic orbits ejected from the Galactic bar or older central regions of the Galactic disk. Little is known about the inner regions of the Galactic disk, whether it exists at all in the inner kpc, what its scale height might be should it exist, etc. The model of \\cite{dirbe_phot}, constructed to reproduce the surface brightness within 5~kpc of the Galactic center seen by COBE/DIRBE at 3.5 and at 4.9~$\\mu$ after correction for extinction, suggests a double exponential disk with a scale heights of 42 and of 210~pc (1.5$^{\\circ}$ at a distance of 8~kpc to the Galactic center) combined with a truncated power-law bulge. However, the beam size of COBE/DIRBE is $0.7^{\\circ} \\times 0.7^{\\circ}$, and this was smoothed to a 1.5$^{\\circ}$ angular resolution for their analysis. The scale height of 210~pc thus corresponds to their minimum angular resolution, and may well be smaller, which would reduce any potential disk contamination of bulge samples. In any case, using their model as a guide, we find that the probability of contamination by disk stars of a sample at $b = 2.5^{\\circ}$ in this maximal case is only $\\sim$30\\% larger than it is at $b = 4^{\\circ}$. Some disk models such as model 2 of \\cite{2mass_bulge}, constructed to fit 2MASS star counts, contain an inner hole, which would lower potential disk contamination of bulge samples substantially. Rather than speculate further on whether or not there is a disk within the central kpc and what its properties might be, we refer back to the large samples of ``bulge'' giants studied in the many references cited above. The innermost field included in most of these is Baade's Window, with $b$ ranging from $-3.9$ to $-4.1^{\\circ}$. The galactic latitudes for the three microlensed dwarfs are $-2.5$, $-3.6$ and $-4.8^{\\circ}$, so that two of these are slightly smaller than that of Baade's window. However, recently \\cite{rich07} presented a detailed abundance analysis from near-IR spectra of M giants in a field at ($l,b$) = ($0^{\\circ},-1^{\\circ}$), significantly closer to the Galactic center than any of the three microlensed dwarfs. Their sample of 17 M giants has the usual properties seen in the Baade's Window samples, a mean [Fe/H] of $-0.22$~dex, with the most metal rich at [Fe/H] $+0.02\\pm0.11$~dex. These giants show the usual $\\alpha$-enhancement. Surely the microlensed dwarfs, each located more than twice as high above the Galactic plane than the giants in this inner bulge field, then cannot be from a different population than the \\cite{rich07} M giant sample; we thus do not believe that the sample of microlensed bulge stars is contaminated by any disk stars that might be co-located near the Galactic center. We suggest that sample bias is responsible for the difference between the median Fe-metallicity seen in the bulge giant samples and that of the microlensed bulge dwarfs. A possible mechanism is the very high mass loss rates predicted to occur at the very high metallicities and moderately high luminosities being discussed here. These are high enough that in an old stellar population, stars are predicted to lose enough mass to peel off the red giant branch (RGB) before reaching the He-flash, which they never go through. Red clump stars, which are burning He in their cores on the horizontal branch, are even more evolved than giants at the RGB tip, and so their evolutionary tracks, once mass loss is properly taken into account, may indicate that their expected numbers in an old very-metal rich population may be even more depleted than are luminous first ascent RGB stars. This effect may have been detected through CMD studies of the extremely metal-rich open cluster NGC~6791 with [Fe/H] +0.45~dex \\citep{carretta07}. In this cluster, \\cite{kalirai07} found a strong relative absence of luminous RGB stars. We propose that this relative paucity of giants on the upper RGB for very metal-rich old populations produces a bias against the highest metallicity stars by preferentially eliminating them from the samples being studied by all previous investigations in the Galactic bulge, which samples consist of one or more of the luminous K giants, luminous M giants, and red clump stars found there. If this is correct, then the mean bulge metallicity may be comparable to that expected from the radial gradients prevailing within the Galactic disk at a time $\\sim$5~Gyr ago extrapolated to the Galactic center. It thus may be that at the present time there is a gradient within the solar circle of stellar metallicity with Galactic radius that is roughly comparable to that measured for the interstellar medium (ISM) at present from HII regions by \\cite{esteban05} and from planetary nebulae (PN) by \\cite{pn_grad}. The latter present an estimate of the change of the radial abundance gradient in the Galaxy as a function of time from PN of varying ages. They suggest that the radial abundance gradient in the ISM in the Galactic disk was twice as large $\\sim$5~Gyr ago when the Sun was formed than it is now. There are important consequences for the chemical evolution of extragalactic objects as well if our conjecture regarding the metallicity distribution of the Galactic bulge is correct. The interpretation of spectra and broad band photometry of the integrated light of simple old metal-rich stellar populations such as are believed to exist in luminous elliptical galaxies may be affected. We expect an underestimate of the true Fe-metallicity of such systems to occur with commonly used tools such as Lick indices, calculated with stellar evolutionary tracks and isochrones that ignore mass loss, as is normally the case. We expect an overestimate of the mass-to-light ratio to occur if the luminous RGB stars we expect to be present near the RGB tip in a metal-rich population are not present in their expected numbers as they contribute only a small fraction of the mass, but a larger fraction of the total luminosity. Changes in the strength of specific absorption features can be expected as well, particularly in the IR, where luminous red giants dominate the integrated light assuming they are present. } We have analyzed a high dispersion spectrum of a microlensed dwarf, \\mystar, in the Galactic bulge. The magnification of this event was very high, HIRES on the 10-m Keck~I Telescope was used, the weather was clear with good seeing, and the exposure time was long compared to any previous such data, so the resulting spectrum has a relatively high signal-to-noise ratio. We stress that in principle the abundance analysis of a upper main sequence dwarf is much easier and less prone to error for spectra of a fixed signal-to-noise ratio than that of a much cooler but much brighter bulge giant with a very complex spectrum full of blends and of strong molecular bands. The advantages of analyzing microlensed bulge stars, for which the required high signal-to-noise ratio can sometimes be achieved, to improving our understanding of the [Fe/H] distribution and chemical evolution of the Galactic bulge are large. We have derived for \\mystar, which we believe to be a dwarf below the main sequence turnoff with \\teff $\\sim 5400$~K, a very high Fe-metallicity, [Fe/H] = \\fehogle ~dex. This is very peculiar given that many previous surveys of the metallicity distribution carried out with large samples of K or M giants in the Galactic bulge find both the mean and median [Fe/H] to be sub-solar. The two other highly magnified Galactic bulge dwarfs studied in detail, \\johnstar\\ by \\cite{johnson07} and \\moa\\ by \\cite{johnson08}, also have very high Fe-metallicities. In order to produce consistency, we suggest that there is a sampling bias in the bulge giant samples such that very metal-rich giants are strongly depleted. We suggest a physical mechanism for this, the very high mass loss rates expected for such metal-rich old giants can exhaust their envelopes prior to the normal He-flash. We also find that \\mystar\\ does not show enhancements of the $\\alpha$ elements; neither does \\johnstar, analyzed by \\cite{johnson07}. However, most bulge giants from samples with high dispersion spectroscopy, e.g. the work of \\cite{fulbright07} and particularly those from \\cite{lecureur07}, do show large (and varying from star to star) enhancements of [Na/Fe], [Mg/Fe] and [Al/Fe] even at super-solar metallicities. We suggest that it is the abundances deduced for the microlensed dwarfs that best represent the initial chemical inventory of the interstellar medium at the time these stars formed, while those derived for the bulge giants may not. We recognize that three stars is a very small sample, but the implications of our results and inferences for the chemical evolution of the Galactic bulge and for the interpretation of integrated light spectra and broad-band photometry of old simple stellar populations such as luminous ellitical galaxies, are so important that we offer these hypotheses at this time. The study of additional highly microlensed Galactic bulge dwarfs to increase the sample from just three such stars is clearly urgent. Now that ongoing microlensing surveys make such observations of Galactic bulge dwarfs feasible, we expect substantial improvements in the sample size of high quality spectra for Galactic bulge dwarfs. Suitable high magnification events are rare and lining up the necessary instruments/telescopes/clear weather at just the right time is difficult. Several years may be required to accumulate a suitable sample of spectra of highly microlensed dwarf stars in the Galactic bulge. We are now appropriately positioned to carry out such a time critical program at the Keck Observatory over the next few years, and eagerly look forward to confirmation of our perhaps premature and provocative hypotheses in the not too distant future." }, "0801/0801.0321_arXiv.txt": { "abstract": "We present detailed radiative transfer spectral synthesis models for the Iron Low Ionization Broad Absorption Line (FeLoBAL) active galactic nuclei (AGN) \\FQBS\\ and \\ISO. Detailed NLTE spectral synthesis with a spherically symmetric outflow reproduces the observed spectra very well across a large wavelength range. While exact spherical symmetry is probably not required, our model fits are of high quality and thus very large covering fractions are strongly implied by our results. We constrain the kinetic energy and mass in the ejecta and discuss their implications on the accretion rate. Our results support the idea that FeLoBALs may be an evolutionary stage in the development of more ``ordinary'' QSOs. ", "introduction": "Spectroscopic observations of quasars show that about 10--20\\% have broad absorption troughs in their rest-frame UV spectra \\citep[see][for example]{trump06}. These absorption lines are almost exclusively blueshifted from the rest wavelength of the associated atomic transition, indicating the presence of an outflowing wind in our line of sight to the nucleus. The line-of-sight velocities range from zero to up to tens of thousands of kilometers per second \\citep[e.g.,][]{narayanan04}. While understanding these outflows is of fundamental interest for understanding the quasar central engine, it is also potentially important for understanding the role of quasars in the Universe. The observation that the black hole mass is correlated with the velocity dispersion of stars in the host galaxy bulge \\citep[e.g.,][]{magorrian98,fm00,gebhardt00} indicates a co-evolution of the galaxy and its central black hole. The close co-evolution implies there must be feedback between the quasar and the host galaxy, even though the sphere of gravitational influence of the black hole is much smaller than the galaxy. Energy arguments, however, show that is quite feasible that the black hole can influence the galaxy; as discussed by \\citet{begelman03}, the accretion energy of the black hole easily exceeds the binding energy of the host galaxy's bulge. The nature of the feedback mechanism that carries the accretion energy to the galaxy is not known. Since AGNs are observed to release matter and kinetic energy into their environment via outflows, it is plausible that these outflows contribute to the feedback in an important way. One of the difficulties in using quasar outflows in this context is that they are sufficiently poorly understood that there are significant uncertainties in such basic properties as the total mass outflow rate and the total kinetic energy. What is the kinetic luminosity of the broad absorption line quasar winds? That turns out to be very difficult to constrain. While the presence of the blueshifted absorption lines unequivocally indicates the presence of high-velocity outflowing gas, the other fundamentally important properties of the gas, including the density, column density, and covering fraction are very difficult to constrain. The density is difficult to constrain because the absorption lines are predominately resonance transitions, and their strengths are not very sensitive to density. Without knowing the density, the distance of the gas from the central engine cannot be constrained; the same ionization state can be attained by dense gas close to the central engine, or rare gas far from the central engine. Density estimates are possible when absorption lines are seen from non-resonance transitions, but even then, they can differ enormously. For example, \\citet{dekool01} analyzed metastable \\ion{Fe}{2} absorption lines in FBQS~0840$+$3633, and inferred a electron density $<1000-3000\\rm\\, cm^{-3}$ and a distance from the central engine of several hundred pc. In contrast, \\citet{eracleous03} analyze the metastable \\ion{Fe}{2} absorption in Arp~102B with photoionization models and infer a density of at least $10^{11}\\rm \\, cm^{-3}$ and a distance of less than $7 \\times 10^{16}\\rm \\, cm$. The global covering fraction is also difficult to constrain directly from the quasar spectrum; we know the gas, at least, partially covers our line of sight, but we have little information about other lines of sight. Covering fraction constraints are generally made based on population statistics. In a seminal paper, \\citet{weymann91} showed that for most BALQSOs, the emission line properties are remarkably similar to non-BALQSOs. Thus, the fact that 10--20\\% of quasar spectra contain broad absorption lines is interpreted as evidence that there is a wind that covers 10--20\\% of sight lines to all similar quasars, and whether or not we see absorption lines depends on our orientation. Alternatively, some BAL quasars have notably different line emission than the average quasar; examples are the low-ionization BALQSOs studied by e.g., \\citet{boroson92b}. These objects may instead represent an evolutionary stage of quasars, as the quasar emerges from the cloud of gas and dust in which it formed \\citep{becker97}. While it seems that the column density should be easy to constrain, more recent work has shown that it can be very difficult to measure. It was originally thought that non-black absorption troughs indicated a relatively low column density for the absorbing gas \\cite[equivalent hydrogen column densities of $10^{19-20}\\rm \\,cm^{-2}$, e.g.,][]{hamann98}. But it has now been found that the non-black troughs indicate velocity-dependent partial covering, where the absorption covers part of the emission region, and the uncovered part fills in the trough partially \\citep[e.g.,][]{arav99}. Thus, the column density appears to be high, but it is very difficult to constrain directly from the data except in a few very specialized cases \\citep[see for example][]{gabel06,arav05}. How can we make progress on this problem? It is becoming clear that because of the difficulties described above, the traditional techniques for analysis of troughs (e.g., curve of growth) and modeling (e.g., photoionization modeling to produce absorption line ratios and equivalent widths) are limited. An approach that may be profitable is to construct a physical model for the outflow, and constrain the parameters of the model using the data. Our first foray into constructing physical models for quasar winds was performed by \\citet{branch02}. In that paper, the FeLoBAL\\footnote{FeLoBALs are distinguished by the presence of absorption in low-ionization lines such as \\ion{Al}{3} and \\ion{Mg}{2} as well as absorption by excited states of \\ion{Fe}{2} and \\ion{Fe}{3}.} FIRST~J121442$+$280329 was modeled using SYNOW, a parameterized, spherically-symmetric, resonant-scattering, synthetic spectrum code more typically used to model supernovae \\citep{fisher00}. The difference between this treatment and a more typical one applied to the same data by \\citet{dekool02b} is that SYNOW assumes that emission and absorption are produced in the same outflowing gas. In contrast, the approach taken by \\citet{dekool02b} assumes that absorption is imprinted upon a typical continuum+emission line quasar spectrum; that is, the absorbing gas is separated from the emission-line region. In fact, based on the analysis of the \\ion{Fe}{2} metastable absorption lines, they find that the absorber is 1--30 parsecs from the central engine, much farther than the quasar broad emission-line region. Note that FIRST~J121442$+$280329 is not the only object that can be modeled using SYNOW; \\citet{casebeer04} present a SYNOW model of another FeLoBAL, ISO~J005645.1$-$273816. The SYNOW model is attractive because it is simple; only one component is needed to model both the emission and absorption lines. However, this model is limited. The primary purpose of the SYNOW program is to identify lines in complicated supernova spectra. Thus, individual ions can be added to a SYNOW run at will in order to see if features from emission and absorption from those ions is present. It does not solve the physics of the gas, so physical parameters beyond the existence of a particular species and its velocity extent cannot be extracted from the results. We test the ideas of \\citet{branch02} and \\citet{casebeer04} by using the generalized stellar atmosphere code \\texttt{PHOENIX} to model the spectra of the two FeLoBALs that were successfully modeled using SYNOW, and including spectra that extend to rest-frame optical wavelengths for FIRST~J121442$+$280329. {\\tt PHOENIX} is a much different code than \\texttt{SYNOW} in that it contains all the relevant physics to determine the spectrum of outflowing gas. It solves the fully relativistic NLTE radiative transfer problem including the effects of both lines and continua in moving flows. For a discussion of the use of both \\texttt{SYNOW} and \\texttt{PHOENIX} in the context of modeling supernovae spectra, see \\citet*{bbj03}. We find that \\texttt{PHOENIX} is able to model the spectra from these objects surprisingly well, and we are able to derive several important physical parameters from the model. In \\S \\ref{Models} we describe the \\texttt{PHOENIX} model in detail. In \\S \\ref{Modelfits} we describe our determination of the best-fitting model. In \\S \\ref{Results} we describe the results of our model fitting. In \\S \\ref{Discussion} we discuss the physical implications of the model, how it relates to other BAL spectra and where it fits in the BAL picture. An appendix includes a flowchart of a \\phx\\ calculation. ", "conclusions": "Using the spectral synthesis code \\texttt{PHOENIX} we compute synthetic spectra that provide very good fits to the observed restframe UV and optical spectra of two FeLoBALs. While our models are limited to exact spherical symmetry, they provide excellent fits. In order to reconcile our results with the polarization data on these objects, we require some asymmetry, but still a high global covering fraction, which would only modestly affect the flux spectrum. We are able to determine a luminosity distance estimate which is direct and is accurate to around 50\\%. The question arises: could these objects be used as distance indicators at high z, even if only as a sanity check on the really high-z Hubble diagram from GRBs? Our results lend support to the inference that FeLoBALs are an evolutionary stage of the QSO as opposed to a pure orientation effect. Our model with a smaller covering factor may be able to explain other BAL QSO such as overlapping trough QSOs. In addition our model column densities, which are Compton thick, match those that are expected from X-ray non-detections of these objects. For future work we plan to explore metallicity effects on the model spectrum. This is not trivial, as it is not just a matter of changing the metallicity and comparing the model. Completely different sets of dynamical and luminosity parameters may be required to achieve the best fitting model spectrum. We plan to continue our analysis with the \\citet{hall07} spectrum, which shows $H\\beta$ absorption. We have compared our observed spectrum with the one from that paper and think that it is an excellent candidate for this type of modeling." }, "0801/0801.2923_arXiv.txt": { "abstract": "The sources of nuclear uncertainties in nova nucleosynthesis have been identified using hydrodynamical nova models. Experimental efforts have followed and significantly reduced those uncertainties. This is important for the evaluation of nova contribution to galactic chemical evolution, gamma--ray astronomy and possibly presolar grain studies. In particular, estimations of expected gamma--ray fluxes are essential for the planning of observations with existing or future satellites. ", "introduction": "Novae are thermonuclear runaways occurring at the surface of a white dwarf accreting hydrogen rich matter from its companion in a close binary system\\cite{Sta72,Geh98,JH98,JH07}. Material from the white dwarf $^{12}$C and $^{16}$O (CO nova) or $^{16}$O, $^{20}$Ne plus some Na, Mg and Al isotopes (ONe nova) provide the seeds for the operation of the CNO cycle and further nucleosynthesis. Novae are supposed to be at the origin of galactic $^{15}$N and $^{17}$O and contribute to the galactic chemical evolution of $^7$Li and $^{13}$C. In addition they produce radioactive isotopes that could be detected by their gamma--ray emission: $^7$Be (478~keV), $^{18}$F ($\\le$511~keV), $^{22}$Na (1.275~MeV) and $^{26}$Al (1.809~MeV). The yields of these isotopes depend strongly on the hydrodynamics of the explosion but also on nuclear reaction rates involving stable and radioactive nuclei. Tests of sensitivity to the reaction rates uncertainties have been done using parametrized\\cite{Wor94}, semi--analytic\\cite{Coc95}, post--processed\\cite{Ili02} nova models but also with a 1-D hydrodynamical model. Indeed, in a series of papers the impact of nuclear uncertainties in the hot-pp chain\\cite{Be96}, the hot-CNO cycle\\cite{F00}, the Na--Mg--Al region\\cite{NaAl99} and Si--Ar region\\cite{SiAr01} have been investigated with the Barcelona (SHIVA) hydrocode. In this way, the temperature and density profiles, their time evolution, and the effect of convection time scale were fully taken into account. The nuclear reaction rates whose uncertainties affected most nova nucleosynthesis having been identified, many nuclear physics experiments were conducted to reduce these uncertainties. In this review, we will shortly summarize the experimental progress made in this domain. ", "conclusions": "Detailed calculations performed with the SHIVA hydrodinamical code have enable the identification of nuclear uncertainties affecting nova nucleosynthesis. We see that less than ten years after, great progress have been made thanks to experimental efforts, in particular for the $^{17}$O(p,$\\gamma)^{18}$F, $^{17}$O(p,$\\alpha)^{14}$N, $^{18}$F(p,$\\gamma)^{19}$Ne, $^{18}$F(p,$\\alpha)^{15}$O, $^{21}$Na(p,$\\gamma)^{22}$Mg, $^{22}$Na(p,$\\gamma)^{23}$Mg $^{25}$Al(p,$\\gamma)^{26}$Si, $^{26g.s.}$Al(p,$\\gamma)^{27}$Si and $^{30}$P(p,$\\gamma)^{31}$S reactions that were identified as the most influential. However, further efforts are required for the $^{22}$Na(p,$\\gamma)^{23}$Mg, $^{25}$Al(p,$\\gamma)^{26}$Si, $^{30}$P(p,$\\gamma)^{31}$S reactions and especially for the $^{18}$F(p,$\\alpha)^{15}$O reaction. For this last reaction where contribution of interfering broad resonance tails are essential, progress should come from direct measurements with intense $^{18}$F beam (TRIUMF) or from indirect (THM\\cite{THM}) measurement planned at the CRIB of the Center for Nuclear Studies (Wako). \\begin{theacknowledgments} I am indebted to Margarita Hernanz and Jordi Jos\\'e for a now more than twelve years collaboration on nova nucleosynthesis and to Nicolas~de~S\\'er\\'eville for frequent discussions. Many thanks also to Carmen Angulo, Christian Iliadis, Fa\\\"{\\i}rouz Hammache, Fran\\c{c}ois de Oliveira Santos and Claudio Spitaleri for long time collaborations. \\end{theacknowledgments}" }, "0801/0801.2937_arXiv.txt": { "abstract": "OH megamasers (OHMs) emit primarily in the main lines at 1667 and 1665 MHz, and differ from their Galactic counterparts due to their immense luminosities, large linewidths and 1667/1665 MHz flux ratios, which are always greater than one. We find that these maser properties result from strong 53 \\mic\\ radiative pumping combined with line overlap effects caused by turbulent linewidths $\\sim$ 20 \\kms; pumping calculations that do not include line overlap are unreliable. A minimum dust temperature of \\about\\ 45 K is needed for inversion, and maximum maser efficiency occurs for dust temperatures \\about\\ 80 -- 140 K. We find that warmer dust can support inversion at lower IR luminosities, in agreement with observations. Our results are in good agreement with a clumpy model of OHMs, with clouds sizes \\la\\ 1 pc and OH column densities \\about\\ 5\\x\\E{16}\\cmsq, that is able to explain both the diffuse and compact emission observed for OHMs. We suggest that {\\em all} OH main line masers may be pumped by far-IR radiation, with the major differences between OHMs and Galactic OH masers caused by differences in linewidth produced by line overlap. Small Galactic maser linewidths tend to produce stronger 1665 MHz emission. The large OHM linewidths lead to inverted ground state transitions having approximately the same excitation temperature, producing 1667/1665 MHz flux ratios greater than one and weak satellite line emission. Finally, the small observed ratio of pumping radiation to dense molecular gas, as traced by HCN and HCO$^+$, is a possible reason for the lack of OH megamaser emission in NGC 6240. ", "introduction": "OH megamasers are extremely luminous extragalactic OH maser sources that have isotropic luminosities a million or more times greater than their Galactic counterparts. OHMs are found only in the nuclear regions of luminous (LIRGs) and ultraluminous (ULIRGs) infrared galaxies, where intense star formation is occurring. These masers emit primarily in the main lines at 1667 and 1665 MHz, although weaker emission in the satellite lines at 1612 and 1720 MHz has also been detected \\citep{baan87}. In addition to their immense luminosities, OHMs differ from their Galactic counterparts in their very large linewidths and their 1667/1665 MHz flux ratios. Galactic OH mainline masers in star forming regions tend to have linewidths \\la\\ 1 \\kms\\ and 1667/1665 MHz flux ratios less than one. In contrast, OHMs have overall linewidths \\ga\\ 100 \\kms\\ and their 1667/1665 MHz flux ratios are always greater than one. Although this ratio varies, the typical ratio is \\about\\ 5 \\citep{lonsdale02}. \\citet{baan89} found a clear separation between OH emitters and absorbers based on the IR properties of the host galaxies. The OHMs have larger IR luminosity (\\LIR) and tend to be warmer. OHMs may occur for lower \\LIR\\ as long as the dust is warm enough. The Arecibo OHM survey conducted by \\citet{darling02} targeted LIRGs and increased the number of OHMs to \\about\\ 100. That survey substantiated the conclusions of \\citet{baan89} and showed that the fraction of OHMs in LIRGs is an increasing function of far-IR (FIR) luminosity and color, with most of the detections found in warm ULIRGs. These observations strongly suggest that the masers are pumped by the intense FIR radiation in the OH pump lines at 53 \\mic\\ and 35 \\mic. Support for FIR pumping is provided by observations using the Infrared Space Observatory (ISO). \\citet{skinner} observed the 35 \\mic\\ pumping transition in Arp 220 and concluded that absorption in this line alone could power its maser emission. Significantly, OH and water megamasers have not been observed in the same galaxy \\citep{lo05} even though both are often found in the same Galactic star forming regions. The standard OHM model was introduced by \\citet{baan85}, who proposed that the maser emission is produced by low gain unsaturated amplification of background radio continuum. This proposal was confirmed by the comprehensive study of OHMs performed by \\citet{henkelwilson90}. They found good agreement with observations if the maser transitions have approximately equal excitation temperatures, giving a 1667/1665 MHz optical depth ratio of \\about \\ 1.8. However, subsequent VLBI observations revealed compact maser emission on parsec scales with amplification factors \\ga \\ 800 and very large 1667/1665 MHz line ratios \\citep{lonsdale98}. A surprise of these observations was that linewidths remained large (tens of \\kms) even on the smallest observed angular scales \\citep{lonsdale02}. The VLBI observations led to the suggestion that there are two different modes of maser operation --- low gain, unsaturated amplification responsible for the diffuse maser emission and high gain, saturated emission producing the compact sources. It was also suggested that the two classes may have different pumping mechanisms, with the diffuse emission pumped by IR radiation and the compact masers by a combination of collisions and radiation \\citep{lonsdale02}. However, an extension of the standard model is able to explain both the diffuse and compact emission from IIIZw35 by assuming a clumpy maser medium \\citep{parra05}. Each cloud generates low gain, unsaturated emission, and strong compact emission occurs when the line of sight intersects more than one cloud. A similar model has been used to explain the megamaser emission from IRAS 17208-0014 \\citep{momjian}. The clumpy maser model also explains the observation that compact masers are always found embedded in the diffuse emission and do not occur in isolation \\citep{lonsdale02}. An observational difficulty is that the nearest OHMs are located at distances of \\about \\ 100 Mpc and thus cannot be probed nearly as accurately as Galactic masers. The study of OHMs thus relies heavily on analyzing their global properties. It is important to first explain the global properties of OHMs and then construct more comprehensive models for those sources which have been observed in more detail using VLBI techniques. Previous OHM models \\citep{henkelwilson90,randell} were performed before the discovery of the compact maser sources. Although the clumpy maser model seems to provide a phenomenological explanation to the compact and diffuse maser emissions, it is not yet backed by pumping calculations. The time is right to develop a pumping model that can explain both the compact and diffuse emission using physical conditions that are consistent with those expected in the maser regions. This is the aim of the present paper. In section 2 we summarize what is known concerning the physical conditions existing in the maser region. In section 3 we explain our model and present the results of our calculations. Section 4 contains a summary and discussion. ", "conclusions": "OHMs are characterized by their extreme luminosity, 1667/1665 MHz flux ratio greater than one, and large linewidths. Our detailed pumping analysis of OHMs shows that the above properties are a natural outcome of an intense FIR pump combined with line overlap effects due to large turbulent linewidths. The 53 \\mic\\ lines provide the primary pump, and a minimum dust temperature of \\about \\ 45 K is needed for main line maser production. In agreement with the observations of \\citet{baan89} and \\citet{darling02}, we find that a smaller \\LIR \\ can support inversion, as long as the temperature is warm enough. We find no conditions where collisions can produce the observed maser emission, and when combined with a strong FIR radiation field, collisions tend to weaken and then thermalize the masers. These results are consistent with the detection of the 183 GHz water maser combined with the lack of detection of the 22 GHz water maser in Arp 220 \\citep{cernicharo}. The 22 GHz maser is collisionally pumped and requires relatively high densities and temperatures for its production, while the 183 GHz maser is produced at much lower densities and temperature. The large OHM linewidths produce significant overlap among the the 53 \\mic\\ pump lines resulting in nearly equal negative excitation temperatures for all ground state transitions, in agreement with observations \\citep[figure \\ref{fig:color-color}; see also][]{henkelwilson90}. The 1667/1665 MHz flux ratios are thus greater than one and only weak satellite line emission is produced. Our pumping results are in agreement with those required by clumpy maser models that explain both the diffuse and compact maser emission from OHMs such as IIIZw35. Finally, the small observed ratios of \\LIR /\\LHCN \\ and \\LIR /\\LHCO+ \\ relative to those found in OHMs is a possible reason for the lack of OH megamaser emission in the ULIRG NGC6240. A minimum ratio of pumping radiation to dense molecular gas may be a prerequisite for becoming a OHM. Excited state transitions at 5 cm and 6 cm have also been observed in OHMs. However, these lines have only been detected in absorption. \\citet{henkel87} observed absorption in the three 6 cm $^2\\Pi_{1/2}(J = 1/2)$ transitions in five OHMs, and absorption has been detected also in one 5 cm $^2\\Pi_{3/2}(J = 5/2)$ transition in Arp 220 \\citep{henkel86}. These observations are surprising since masing in the 5 cm and 6 cm lines has often been observed in conjunction with Galactic ground state OH maser sources \\citep{cesaroni}. Our pumping models of ground state maser emission in OHMs usually predict absorption in the 5 cm lines. However, the 6 cm lines are weakly inverted with opacities a hundred times smaller than the 1667 MHz line. Our calculations find that absorption in the 6 cm lines requires smaller dust temperatures than exist in the maser regions. We expect the absorption to occur outside the maser regions where the dust temperature is lower. Understanding the absorption at 5 cm and 6 cm is an important problem that we plan to treat in a future paper. \\subsection{OH main-line maser pump} Our results suggest that {\\em all} OH main-line masers could be pumped by the same mechanism: far-IR radiation. The major differences between Galactic star forming regions on one hand and evolved stars and OHMs on the other can be attributed to differences in line overlap effects. As shown in \\S\\ref{sec:overlap}, linewidth is the most important factor in determining the 1667/1665 MHz flux ratio and thus is the dominant factor causing the extreme differences between the flux ratios of the two classes. In Galactic star forming regions the observed linewidths are \\la \\ 1 \\kms\\ and the 1665 MHz maser is usually strongest, with the 1667 MHz line often comparable in strength to the satellite lines. In OHMs the linewidths are large (\\ga \\ 10 \\kms) and the 1667 MHz is always stronger than the 1665 MHz line, with the satellite lines much weaker. Our results show that such large linewidths produce approximately equal negative excitation temperatures for the four ground state transitions, resulting in the 1667 MHz line being stronger than the 1665 MHz line and much weaker satellite lines. Smaller linewidths tend to produce a stronger inversion in the 1665 MHz line. In evolved stars, the main-line maser velocities are larger than those found in Galactic star forming regions and they also tend to have 1667/1665 MHz flux ratios \\ga\\ 1. Thus the linewidth could be the main reason why OH main-line masers display differences in spite of sharing a common pump mechanism." }, "0801/0801.3270_arXiv.txt": { "abstract": "We present an investigation into the potential effect of systematics inherent in multi-band wide field surveys on the dark energy equation of state determination for two 3D weak lensing methods. The weak lensing methods are a geometric shear-ratio method and 3D cosmic shear. The analysis here uses an extension of the Fisher matrix framework to jointly include photometric redshift systematics, shear distortion systematics and intrinsic alignments. Using analytic parameterisations of these three primary systematic effects allows an isolation of systematic parameters of particular importance. We show that assuming systematic parameters are fixed, but possibly biased, results in potentially large biases in dark energy parameters. We quantify any potential bias by defining a Bias Figure of Merit. By marginalising over extra systematic parameters such biases are negated at the expense of an increase in the cosmological parameter errors. We show the effect on the dark energy Figure of Merit of marginalising over each systematic parameter individually. We also show the overall reduction in the Figure of Merit due to all three types of systematic effects. Based on some assumption of the likely level of systematic errors, we find that the largest effect on the Figure of Merit comes from uncertainty in the photometric redshift systematic parameters. These can reduce the Figure of Merit by up to a factor of $2$ to $4$ in both 3D weak lensing methods, if no informative prior on the systematic parameters is applied. Shear distortion systematics have a smaller overall effect. Intrinsic alignment effects can reduce the Figure of Merit by up to a further factor of $2$. This, however, is a worst case scenario. By including prior information on systematic parameters the Figure of Merit can be recovered to a large extent, and combined constraints from 3D cosmic shear and shear ratio are robust to systematics. We conclude that, as a rule of thumb, given a realistic current understanding of intrinsic alignments and photometric redshifts, then including all three primary systematic effects reduces the Figure of Merit by \\emph{at most} a factor of $2$. ", "introduction": "\\label{Introduction} It has recently been shown that the equation of state of dark energy could be constrained to a high degree of accuracy using wide and deep imaging surveys (see Albrecht et al., 2006; Peacock et al., 2006; for recent and extensive reviews). 3D weak lensing has been shown to be a particularly powerful way to use the information from such surveys in the determination of dark energy parameters (see Munshi et al., 2006 for a recent review). 3D weak lensing, in which the shear and redshift information of every galaxy is used, has the potential to constrain the dark energy equation of state, $w(z)=\\rho_{\\rm de}(z)/p_{\\rm de}(z)$, to $\\Delta w(z)\\sim 0.01$ using surveys such as Pan-STARRS (Kaiser et al., 2002) and DUNE (Refregier et al., 2006). However, the predictions made thus far (Heavens et al., 2006; Taylor et al., 2007) have only included statistical errors and have not included systematic effects. Since the scientific goal of many future surveys is to constrain $\\Delta w(z)\\sim 0.01$ such systematic effects have the potential to render any cosmological constraints impotent. In this paper we will address astrophysical, instrumental and theoretical systematic effects relevant to multi-band weak lensing surveys in an analytic way. We specifically study the systematic effects of photometric redshifts, intrinsic alignments and shear distortion. We consider these three primary systematics to have potentially the largest effect on the ability of weak lensing surveys to constrain cosmological parameters. Secondary effects, such as source clustering, non-Gaussian effects and theoretical approximations such as the Born approximation have been shown to have a smaller effect on shear statistics (Shapiro \\& Cooray, 2006; Semboloni et al., 2007; Schneider et al., 2002). Note that the effect of non-Gaussianity (Semboloni et al., 2007) has only been studied via simulations, a full analytic investigation could reveal non-Gaussianity to be a larger systematic effect (Takada, private communication). We will examine the degradation that these primary effects may produce in the determination of the dark energy equation of state constraints for two 3D weak lensing methods: 3D cosmic shear (Heavens, 2003; Castro et al., 2005; Heavens et al., 2006) and the shear-ratio method (Jain \\& Taylor, 2003; Taylor et al., 2007). The spirit of the approach taken here is to use simple, analytic, descriptions of systematic effects. By distilling complex effects into simple components the change in cosmological parameter determination due to any particular aspect of a systematic effect can be analysed independently. For example, is the bias or the fraction of outliers in photometric redshifts a more important factor? The obvious penalty in taking such an approach is that the analytic approximations made may not be fully representative of real systematic effects. The two 3D weak lensing methods are introduced in Section \\ref{Methodology}, however we urge the reader to refer to Heavens et al. (2006) and Taylor et al. (2007) for a complete and in-depth introduction to the methods. In Section \\ref{Methodology} we also discuss dark energy parameter prediction. In Section \\ref{Primary Systematic Effects} we introduce the primary systematic effects considered, and the parameterisations used. The potential bias in dark energy parameters due to each systematic effect is presented in Section \\ref{Bias in Dark Energy Parameters}. A marginalisation over systematic parameters in presented in Section \\ref{Marginalising over systematic parameters}. We conclude and recap with a discussion in Section \\ref{Discussion}. For any technical details concerning the 3D weak lensing methods and systematic parameters see Appendix A and B. ", "conclusions": "\\label{Conclusion} We have shown using simple analytic approximations that systematic effects can have a substantial impact on the ability of 3D weak lensing methods, the shear-ratio method and the 3D cosmic shear method, to constrain the dark energy equation of state. We used the Figure of Merit (FoM) to gauge the ability of a next generation experiment to constrain the dark energy equation of state. The systematic effects we considered are those associated with photometric redshifts, an overall distortion in the image plane and intrinsic alignments (both the GI and II terms). The dark energy FoM can be degraded a factor of $2$ due to the photometric redshift systematics alone, for both 3D weak lensing methods. Shear distortion systematics have a small effect for both methods. Intrinsic alignment effects alone can degrade the FoM by a further factor of $2$ for the 3D cosmic shear method, but have a small effect for the shear-ratio method. This difference is due to the way in which the methods use shear information. When an extra systematic parameter is encountered it can either be marginalised over using the available data (equivalent to self-calibration), thereby increasing the marginal error on any cosmological parameter of interest, or it can be assumed to be fixed. If a parameter is fixed then any deviation away from the assumed value will bias the most likely value of any measured cosmological parameter. This bias is a function of both the cosmological parameter error and the sensitivity of a method to any systematic parameter. From this analysis it has been shown that assuming some parameters to be fixed can lead to large biases in $w(z_p)$ and $w_a$, this is complimentary to the analysis done by Amara \\& Refregier (2007) who investiagted the bias in cosmological parameters due to shear measurement systematics using weak lensing tomography. The methods are remarkably insensitive to many parameters, in particular most of the photometric redshift parameters, including the fraction of outliers, and an overall distortion of the shear field. By adopting a parameterisation of the photometric-spectroscopic plane we have shown that a bias in any photometric redshift redshift technique needs to be known to within $\\pm 10^{-3}$ for the dark energy FoM to remain unaffected. We have shown that to calibrate the photometric redshift parameters approximately $10^5$ spectroscopic redshifts are required for the shear-ratio method and approximately $10^2$ to $10^3$ for the 3D cosmic shear method. This difference can be attributed to the binning in redshift required by the shear-ratio method, which may also explain the agreement between the predicted spectroscopic requirements of the shear-ratio and shear tomography methods. The intrinsic alignment terms were modelled using the Heymans et al. (2006) analytic approximations. The GI and II terms had a small effect on the FoM from the shear-ratio method, we found a drop in the FoM of approximately $10\\%$ when the extra covariances were included. For 3D cosmic shear the FoM is reduced by approximately $50\\%$, but there is a very small sensitivity to the extra intrinsic alignment systematic parameters. The 3D cosmic shear result is in agreement to what has been found using shear tomography in Bridle \\& King (2007). The caveat to these comparisons is that we use a fully 3D cosmic shear method, with no binning, and only investigate the Heymans et al. (2006) parameterisation; Bridle \\& King (2007) consider a variety of parameterisations and investigate a binning tomographic method. We have shown that a good prior on systematic parameters can improve the FoM. The relative reduction in the FoM can be limited to $\\leq 30\\%$ if the prior on all systematic parameters has a Gaussian error of $\\sigma_P=0.001$. In particular we have shown that a prior on $z_{\\rm bias}$ can improve the FoM. Good priors on $z_{\\rm bias}$, $c_{\\rm calibration}$ would be particularly helpful in limiting the effect on the FoM. By combining the 3D weak lensing methods the FoM can be increased by a up to a factor of $6$ relative to the methods individually. Furthermore the photometric redshift systematic parameter degeneracies are complementary leading to less systematic degradation in the combined constraints. The bottom line is that the most important systematics to control are those concerning the photometric redshift distribution. If these can be controlled to 1\\% accuracy, then the FoM for proposed future surveys such as DUNE and Pan-STARRS may be reduced by at most a factor of order two. In order to reduce these systematics to a negligible level needs $0.001$ accuracy in median redshifts, requiring ${\\mathcal O} (10^4)$ redshifts. We make a number of recommedations and observations with which to guide future systematic investigations \\begin{itemize} \\item Photometric redshift systematics play a dominant role in the systematics that affect 3D weak lensing. The individual systematic parameter which can have the largest effect on the FoM is the bias in photometric redshifts. However approximately ${\\mathcal O} (10^4)$ spectroscopic redshifts should be enough to calibrate photometric redshifts to the required accuracy. These would need to be representative of the photometric galaxies and complete. \\item If shear calibration bias is assumed to be fixed then an uncertainty in the bias of $\\sim 0.008$ can bias dark energy parameters by $>0.01$. However marginalising over shear bias has a smaller effect on the FoM. \\item Intrinsic alignments play a major role in 3D weak lensing systematic effects, and can reduce the maximum achievable FoM by up to $4$. The broad agreement between the parameterisations investigated here and in Bridle \\& King (2007) suggest that the general trends are robust. \\end{itemize} Despite degrading systematic effects 3D weak lensing retains the potential to be the most powerful cosmological probe of dark energy. Weak lensing is entering a formative period in its development, given that the statistical ability of the method to constrain cosmology is accepted attention must now be focussed on understanding and reducing systematic effects." }, "0801/0801.3931_arXiv.txt": { "abstract": "\\label{abst} We study the dynamics of 2D and 3D barred galaxy analytical models, focusing on the distinction between regular and chaotic orbits with the help of the Smaller ALigment Index (SALI), a very powerful tool for this kind of problems. We present briefly the method and we calculate the fraction of chaotic and regular orbits in several cases. In the 2D model, taking initial conditions on a Poincar\\'{e} $(y,p_y)$ surface of section, we determine the fraction of regular and chaotic orbits. In the 3D model, choosing initial conditions on a cartesian grid in a region of the $(x, z, p_y)$ space, which in coordinate space covers the inner disc, we find how the fraction of regular orbits changes as a function of the Jacobi constant. Finally, we outline that regions near the $(x,y)$ plane are populated mainly by regular orbits. The same is true for regions that lie either near to the galactic center, or at larger relatively distances from it. ", "introduction": "\\label{Intro} The dynamical evolution of galactic systems depends crucially on their orbital structure and in particular on what fraction of their orbits is regular, or chaotic. Thus to permit further studies, it is essential to be able to distinguish between these two types of orbits in a manner that is both safe and efficient. This is not trivial and becomes yet more complicated in systems of many degrees of freedom. A summary of the methods that have been developed over the years can be found in \\cite{Con_spr}. In the present paper we use a method based on the properties of two deviation vectors of an orbit, the ``Smaller ALingment Index\" (SALI) \\cite{sk:1}. It has been applied successfully in different dynamical systems \\cite{sk:1,sk:3,sk:5,Pan:1,Ant:2,Ant:3,Bou:1,Man:1,Man:2,Man:3,Man:4}, frequently also under the name Alignment Index (AI) \\cite{VKS1,VKS2,VHC,KVC,KEV} and has been shown to be a fast and easy to compute indicator of the chaotic or ordered nature of orbits. We first recall its definition and we then show its effectiveness in distinguishing between ordered and chaotic motion by applying it to a barred potential of 2 and 3 degrees of freedom. Recently, a generalization of the SALI, the ``Generalized ALignment Index\" (GALI) was introduced by Skokos et al. (2007) \\cite{sk:6}, which includes the full set of the $k$ initially linearly independent deviation vectors of the system to determine if an orbit is chaotic or not. \\vspace{-0.5 cm} ", "conclusions": "\\label{Conclusions} We used SALI to study the dynamical behavior of Hamiltonian models of 2D and 3D barred galaxies. We found that in both cases there is a significant amount of chaotic orbits. In the 2D model, we were able to chart a subspace of the phase space, to identify rapidly even tiny regions of regular motion and measure their percentages. In the 3D model, apart from computing the global percentages of regular and chaotic orbits, we calculated these percentages as a function of the energy and found that low values of the energy are mainly dominated by `regular' orbital motion. We also followed the distribution of the chaotic and regular orbits in the configuration space and we found that orbits which lie near the $(x,y)$--plane with relatively small mean deviations in $z$--direction are generally regular. Finally, we monitored the variation of their percentages as a function of their initial spherical radius and their mean spherical radius. We find that the fraction of regular orbits is dominant in regions near the center, as well as at relatively larger distances from it. \\vspace{-0.5 cm}" }, "0801/0801.1105_arXiv.txt": { "abstract": "We investigate the properties of one--dimensional flux ``voids'' (connected regions in the flux distribution above the mean flux level) by comparing hydrodynamical simulations of large cosmological volumes with a set of observed high--resolution spectra at $z\\sim 2$. After addressing the effects of box size and resolution, we study how the void distribution changes when the most significant cosmological and astrophysical parameters are varied. We find that the void distribution in the flux is in excellent agreement with predictions of the standard $\\Lambda$CDM cosmology, which also fits other flux statistics remarkably well. We then model the relation between flux voids and the corresponding one--dimensional gas density field along the line--of--sight and make a preliminary attempt to connect the one--dimensional properties of the gas density field to the three--dimensional dark matter distribution at the same redshift. This provides a framework that allows statistical interpretations of the void population at high redshift using observed quasar spectra, and eventually it will enable linking the void properties of the high--redshift universe with those at lower redshifts, which are better known. ", "introduction": "Over the past decade, properties of voids have been more widely investigated, using different observational probes and tracers (mainly in the local universe, see for example \\cite{hoyle}, \\cite{rojas}, \\cite{goldberg}, \\cite{rojas05}, \\cite{hoyle05}, \\cite{ceccarelli}, \\cite{patiri06a}, \\cite{patiri06b}, \\cite{tikhonov}), and theoretical -- analytical or numerical -- models in the framework of the $\\Lambda$ Cold Dark Matter ($\\Lambda$CDM) concordance cosmology (e.g. \\cite{peebles}, \\cite{arbabi}, \\cite{mathis}, \\cite{benson}, \\cite{gottlober}, \\cite{sheth}, \\cite{goldberg04}, \\cite{bolejko}, \\cite{colberg}, \\cite{padilla}, \\cite{furlanettopiran}, \\cite{hoeft}, \\cite{leepark}, \\cite{patiri06c}, \\cite{shandarin}, \\cite{aloisio}, \\cite{tully}, \\cite{parklee}, \\cite{brunino}, \\cite{vdw07}, \\cite{neyrinck}, \\cite{peebles07}). The emerging picture is encouraging. On the one hand, some results appear to be somewhat hard to understand, such as the fact that galaxies observed at the edges of and in voids appear to be a fair sample of the whole galaxy population (the so-called 'void phenomenon', \\cite{peebles}), the fact that dwarf galaxies are not found in voids contrary to expectations from numerical simulations (e.g. \\cite{peebles07}, or the cold spot observed in the Cosmic Microwave Background (CMB) data (\\cite{rudnick}, \\cite{naselsky}). On the other hand, other observational results, such as those from the 2dF or SDSS galaxy redshift surveys, from studies of voids in quasar spectra, or the CMB spectrum (\\cite{caldwell}), are in reasonably good agreement with theoretical predictions from numerical simulations of the standard $\\Lambda$CDM cosmology. Linking the low--redshift properties of voids to those at higher redshifts offers the opportunity to constrain the void population over a significant fraction of cosmic time and to possibly find out more about voids, to ultimately understand their role in the context of cosmic structure formation. Unfortunately, there are very few observables that constrain the population of voids at high redshift (see, however, \\cite{aloisio}). Here, we will use the \\lya forest as a tracer of voids in the high--redshift universe. The \\lya forest (e.g. \\cite{bi}) has been shown to arise from the neutral hydrogen embedded in the mildly non--linear density fluctuations around mean density in the Intergalactic Medium (IGM), which faithfully traces the underlying dark--matter distribution at scales above the Jeans length. It is a powerful cosmological tool in the sense that it probes the dark matter power spectrum in a range of scales ($1$ to $80\\,h^{-1}$\\,Mpc comoving) and redshifts ($z = $2--5.5) not probed by other observables. However, the observable is not the matter distribution itself but the flux: a one--dimensional quantity sensitive not only to cosmological but also to the astrophysical parameters that describe the IGM. The so--called fluctuating Gunn--Peterson approximation \\citep{gunnpeterson} relates the observed flux to the IGM overdensity in a simple manner once the effects of peculiar velocities, thermal broadening, and noise properties are neglected: \\begin{equation} F({\\bf{x}},z)=\\exp[-A(z)(1+\\delta_{IGM}({\\bf{x}},z))^{\\beta}]\\, , \\end{equation} with $\\beta(z) \\sim 2-0.7(\\gamma(z)-1)$, $\\gamma$ the parameter describing the power--law ('equation of state') relation of the IGM, $T=T_0(z)((1+\\delta_{IGM})^{\\gamma-1}$, and $A(z)$ a constant of order unity which depends on the mean flux level, on the Ultra Violet Background (UVB) and on the IGM temperature at mean density. In practice, linking the flux to the density of the dark (or gaseous) matter requires the use of accurate hydrodynamical simulations that incorporate the relevant physical and dynamical processes. However, for approximate calculations equation\\,(1), in conjunction with simple physical prescriptions of \\lya clouds (e.g. \\cite{schaye}), can offer precious insights. Provided the flux--density relation is accurately modelled, using the \\lya forest to check the properties of the population of voids in the matter distribution would avoid having to know the bias between galaxies and dark matter and would allow to sample over larger volumes the values of cosmological parameters that affect the void population. On the other hand, the fact that the flux information is one--dimensional requires non--trivial efforts to relate it to the three--dimensional (dark) matter density field. Motivated by huge amounts of new data of exquisite quality, we here present a first attempt in this direction. Note that in the early 1990s, before the new paradigm of the \\lya forest absorption was proposed, voids from the transmitted \\lya flux of QSO spectra were analysed, using discrete statistics on the lines, by several groups (e.g. \\cite{crotts,duncan,ostriker,dobrzycki,rauch92}). These attempts, however, did not reach conclusive results due to small--number statistics. We defer to a future paper for a link between the observational properties outlined here and those at low redshift (e.g. \\cite{mathis,hoyle}) and for a careful investigation of the absorber--galaxy relation (e.g. \\cite{stocke,mclin}). This paper is organized as follows. In Section \\ref{data}, we briefly describe the data set used. In Section \\ref{simulations} we present the hydrodynamical simulations. Section \\ref{results} contains the bulk of our analysis and the results, which are summarized in the conclusions in Section \\ref{conclusions}. ", "conclusions": "\\label{conclusions} We used \\lya forest QSO spectra to constrain the void population at $z\\sim 2$ at a range of scales and redshifts, which cannot be probed by other observables. The main conclusions can be summarized as follows: \\begin{itemize} \\item the properties (sizes) of voids identified in simulations from the flux distribution as connected regions above the mean flux level are in good agreement with observed ones; \\item using a set of hydrodynamical simulations and varying all the astrophysical and cosmological parameters, along with checking box size and resolution effects, we find that the flux void size distribution is a robust statistics that depends primarily on the flux threshold chosen to define the voids; \\item the flux seems to be a reliable tracer of the underlying dark matter distribution, altering the statistical properties of the dark matter density field (changing amplitude, shape, adding or suppressing power in a wavenumber dependent manner) changes the properties of the void population although at present is difficult to give constraints on cosmological parameters using these statistics; \\item flux voids, i.e. connected regions above the mean flux levels, correspond to gas densities around the mean density, regardless of the role of peculiar velocities, which contribute to a scatter in this relation, but do not alter significantly the mean values; \\item linking 1D voids to the corresponding 3D dark matter voids is more difficult, and we used a void--finding algorithm for that: we find that 3D DM voids with a mean overdensity of $\\delta_{t}=-0.5$ correspond to the flux voids that we have defined from the QSO spectra. However, the correspondence is good only for voids with sizes larger than about $7-10$\\,Mpc$/h$. \\end{itemize} In this paper, we made a preliminary attempt to link the population of voids in the transmitted \\lya flux to the underlying 1D gas density and temperature and 3D dark matter density. The use of \\lya high--resolution spectra is important in the sense that it explores a new regime in scales, redshifts and densities which is currently not probed by other observables: the scales are of order few to tens of Mpc, the redshift range is between $z=2-4$, while the densities are around the mean density. Further studies, that could possibly rely on wider data sets such as the low--resolution SDSS data (\\cite{shang}), on other future observables at higher redshifts (\\cite{aloisio}) or on tomographic studies involving QSO pairs or multiple line--of--sights (\\cite{dodorico}, \\cite{saitta07}) would be important in understanding the dynamical, thermal and chemical evolution of the void population and the interplay between galaxy and the IGM over a large fraction of the Hubble time." }, "0801/0801.4516_arXiv.txt": { "abstract": "We study the anisotropy signature which is expected if the sources of ultra high energy, $>10^{19}$~eV, cosmic-rays (UHECRs) are extragalactic and trace the large scale distribution of luminous matter. Using the PSCz galaxy catalog as a tracer of the large scale structure (LSS) we derive the expected all sky angular distribution of the UHECR intensity. We define a statistic, that measures the correlation between the predicted and observed UHECR arrival direction distributions, and show that it is more sensitive to the expected anisotropy signature than the power spectrum and the two point correlation function. The distribution of the correlation statistic is not sensitive to the unknown redshift evolution of UHECR source density and to the unknown strength and structure of inter-galactic magnetic fields. We show, using this statistic, that recently published $>5.7\\times10^{19}$~eV Auger data are inconsistent with isotropy at $\\simeq98\\%$ CL, and consistent with a source distribution that traces LSS, with some preference to a source distribution that is biased with respect to the galaxy distribution. The anisotropy signature should be detectable also at lower energy, $>4\\times10^{19}$~eV. A few fold increase of the Auger exposure is likely to increase the significance to $>99\\%$ CL, but not to $>99.9\\%$ CL (unless the UHECR source density is comparable or larger than that of galaxies). In order to distinguish between different bias models, the systematic uncertainty in the absolute energy calibration of the experiments should be reduced to well below the current $\\simeq25\\%$. ", "introduction": "The origin of UHECRs is still unknown: Sources have not been identified by cosmic-ray observations and all models of particle acceleration in known astrophysical objects are challenged by the extension of the spectrum to energies exceeding $10^{20}$~eV \\cite{Bhattacharjee:1998qc,Waxman_CR_rev}. The spectrum flattens (becomes harder) \\cite{Nagano:2000ve} and there is evidence for a composition change from heavy to light nuclei \\cite{Bird93,Abbasi05} at $\\sim10^{19}$~eV. This suggests that while the cosmic-ray flux below $\\sim 10^{19}$~eV is dominated by Galactic sources of heavy nuclei, it is dominated at higher energy by different sources of lighter nuclei. The isotropy of the UHECR arrival direction distribution suggests that these sources are extra-Galactic. If UHECRs are produced by a population of extra-Galactic astrophysical objects, that traces the distribution of luminous matter, the inhomogeneous distribution of luminous matter is expected to imprint a characteristic anisotropy signature on the UHECR arrival direction distribution. If UHECRs are indeed light nuclei, this anisotropy is expected to be significant above $\\sim5\\times10^{19}$~eV, where the propagation distance of light cosmic-ray nuclei is limited by interaction with the background radiation field to a distance of the order of 10's of Mpc (\\cite{Greisen:1966jv,Zatsepin:1966jv,Stecker:1999}, see \\cite{Sigl01} for a detailed review). This distance is comparable to that over which order unity variations in the density distribution of luminous matter are observed. Identification of the expected anisotropy signature would provide strong support to models where UHECRs are accelerated in known astrophysical objects, and would be inconsistent with most new physics, \"top-down,\" models, in which UHECRs are produced by the decay of heavy relic particles or topological defects \\cite{Bhattacharjee:1998qc}. The expected anisotropy signal was analyzed in \\cite{Waxman:1996hp}. It was shown there that the signal should be detectable with high statistical significance once the number of detected UHECR events is increased beyond that available at that time, which was $\\approx20$ events above $4\\times10^{19}$~eV, by a factor $\\gtrsim10$. In this paper we revisit this subject, as the commissioning of the Auger detector may provide us within a few years with the required exposure \\cite{Yamamoto:2007xj}. While the analysis of the current paper largely follows that of \\cite{Waxman:1996hp}, it is improved in several respects. First, for the derivation of the large-scale structure of luminous matter we use a larger, as well as more complete and uniform, galaxy redshift survey, the PSCz catalog \\cite{Saunders:2000af} instead of the IRAS 1.2~Jy catalog \\cite{Fisher:1995kd}. The PSCz catalog contains 3 times more galaxies, including many more galaxies at distances of $150-300$ Mpc. Second, we use an updated cosmological model, $\\{\\Omega_m=0.27,\\ \\Omega_\\Lambda=0.73,\\ H_0=75\\ {\\rm km/s/Mpc}\\}$ instead of $\\{\\Omega_m=1,\\ \\Omega_\\Lambda=0,\\ H_0=100\\ {\\rm km/s/Mpc}\\}$, and analyze the effects of a possible redshift evolution of the UHECR source density. Third, we study the sensitivity of the correlation statistic to systematic errors in the determination of UHECR energies. Finally, we compare our proposed method for identifying the expected anisotropy signature to the more commonly used methods based on the power spectrum (e.g.~\\cite{Sommers:2000us,Deligny:2004dj,Anchordoqui:2003bx,Sigl}) and on the two point correlation function (e.g.~\\cite{Takeda:1999sg,DeMarco:2006a,Cuoco:2007id,Takami:2007vv} and references therein) of the angular distribution of UHECR arrival directions. In order to facilitate the use of the proposed statistic for the detection of anisotropy in UHECR data, we have made available at \\verb\"http://www.weizmann.ac.il/\"$\\sim$\\verb\"waxman/criso\" a numerical description of the intensity maps (see \\S~\\ref{sec:summary}). An analysis of the expected anisotropy signal based on the PSCz catalog has recently been carried out in \\cite{Cuoco:2006}. The most important effects taken into account in the present analysis, and neglected in ref.~\\cite{Cuoco:2006}, are the dependence of the anisotropy signal on the finite number density of sources and on the systematic errors in the determination of UHECR energies. As shown in \\S~\\ref{sec:results} and discussed in \\S~\\ref{sec:summary}, both the finite number density of sources and the systematic uncertainties in energy determination have important implications to the predicted signal and its interpretation. We assume in our analysis that the UHERCs are protons. This assumption is motivated by two arguments. First, the observed spectrum of $>10^{19}$~eV cosmic-rays is consistent with a cosmological distribution of proton accelerators producing (intrinsically) a power-law spectrum of high energy protons, $d\\log n/d\\log E\\approx -2$, \\cite{Waxman:1995dg,Bahcall:2002wi} (see also fig.~\\ref{fig:spectrum}). This intrinsic power-law spectrum is consistent with that expected in most models of particle acceleration (\\cite{Waxman_CR_rev} and references therein). Second, the leading candidate extra-Galactic sources (GRBs and AGN, \\cite{Waxman_CR_rev} and references therein) are expected to accelerate primarily protons. The paper is organized as follows. In sec.~\\ref{sec:Method} we describe the formalism used to analyze the anisotropy signal and present maps of the predicted angular distribution of UHECR intensity. In sec.~\\ref{sec:results} we study the sensitivity of several statistics to the predicted anisotropy signature. In sec.~\\ref{sec:Auger} we analyze Auger data, which was published during the preparation of this manuscript \\cite{AugerSc:2007}, using our correlation statistic. Sec.~\\ref{sec:summary} contains a brief summary of the results and a discussion of their implications. The following point should be made here regarding the isotropy analysis of ref.~\\cite{AugerSc:2007}. In ref.~\\cite{AugerSc:2007} a correlation is found between the arrival direction distribution of $>5.7\\times10^{19}$~eV cosmic-rays (detected by the Auger experiment) and between the angular distribution of low-luminosity AGNs included in the V-C AGN catalog \\cite{V-C}. According to the analysis of ref.~\\cite{AugerSc:2007}, the probability that the detected correlation would arise by chance for an isotropic UHECR arrival direction distribution is $\\simeq1$\\%. Since low luminosity AGNs trace the distribution of luminous matter, one may argue that the detected correlation provides the first evidence for a correlation of UHECR sources and LSS. Unfortunately, the V-C catalog is merely a compilation of AGN data available in the literature, and is therefore incomplete both in its sky coverage and in its luminosity coverage. It does not, therefore, provide a correct description of the local LSS, as clearly pointed out in the introduction of ref.~\\cite{V-C}: \"The V-C catalog should not be used for any statistical analysis as it is not complete in any sense, except that it is, we hope, a complete survey of the literature\". The interpretation of the correlation reported in ref.~\\cite{AugerSc:2007} is thus unclear. ", "conclusions": "\\label{sec:summary} We have derived the expected angular dependence of the UHECR intensity, under the assumption that the UHECRs are protons produced by extra-Galactic sources that trace the large scale distribution of luminous matter. All sky maps of the UHECR intensity, $I(E,\\hat\\Omega)$ with several energy thresholds $E$, are presented in figure~\\ref{fig:CRmap1} for an unbiased UHECR source distribution, where the UHECR source density is proportional to the LSS density, and in fig.~\\ref{fig:CRmap2} for a biased model, where the UHECR source density is proportional to the LSS density in over-dense region and vanishes elsewhere. A numerical representation of the intensity maps may be downloaded from \\verb\"http://www.weizmann.ac.il/\"$\\sim$\\verb\"waxman/criso\". The angular structure of the UHECR intensity map reflects the local large scale structure of the galaxy distribution, as can be seen by comparing figures~\\ref{fig:CRmap1} and~\\ref{fig:CRmap2} to figure~\\ref{fig:densityMap75}, which presents the integrated galaxy density out to a distance of 75~Mpc. The anisotropy is larger for higher energy threshold, since the propagation distance increases at lower energy (see fig.~\\ref{fig:DiffFlux}) and the contribution to the UHECR flux of sources beyond $\\sim100$~Mpc constitutes a roughly isotropic \"background\". We have used simple analytic arguments (\\S~\\ref{sec:B}) to show that inter-galactic magnetic fields may modify the intensity map on scales of a few degrees, but not on larger scales (a result which is consistent with detailed semi-analytic and numeric analyses \\cite{Dolag:2004kp,Lemoine08}). We have defined a statistic, $X_C$ (eq.~\\ref{eq:X_C}), that measures the correlation between the predicted and observed UHECR arrival direction distributions, and showed that it is more sensitive to the expected anisotropy signature than the power spectrum and the two point correlation function (see table~\\ref{table:comp}). The value of $X_C$ for a given data set of UHECR arrival directions, the average value of $X_C$ over realizations of the various models (isotropic, I, un-biased, UB, biased, B), and the distributions of $X_C$ values expected in realizations of the various models in the limit of infinite UHECR source density, can all be straightforwardly calculated using the numerical representations of the UHECR maps at \\verb\"http://www.weizmann.ac.il/\"$\\sim$\\verb\"waxman/criso\". In order to avoid sensitivity to possible distortions of the UHECR intensity map by magnetic fields, we have used $6^\\circ\\times6^\\circ$ bins in calculating $X_C$ and excluded the Galactic plane region, $|b|<12^\\circ$. As can be seen by comparing results in tables~\\ref{table:comp} and~\\ref{table:opt}, the exclusion of the Galactic plane region does not affect the results significantly. Table~\\ref{table:opt} demonstrates that the anisotropy signal is stronger at lower energy: Although the contrast of the fluctuations in the UHECR intensity is higher at high energy (see fig.~\\ref{fig:CRmap1}), the signal becomes weaker at higher energies since the number of observed UHECR drops rapidly with energy. At $E\\sim20$~EeV a significant contamination from Galactic sources may be present and the deflection of cosmic-ray particles may become significant. In order to avoid distortions of the UHECR intensity map by these effects, it is advisable to choose an energy threshold of $E>40$~EeV. We have shown that the distribution of $X_C$ is not sensitive to the assumed redshift evolution of the source density distribution, to variations in the intrinsic spectrum of protons produced by the sources, and to statistical errors in the experimental determination of event energies (see fig.~\\ref{fig:sens} and table~\\ref{table:opt}). The distribution of $X_C$ does depend on the assumed local ($z=0$) source number density, $\\bar{s}_0$: The average value of $X_C$ obtained in realizations of the various models (I, UB, B) is independent of $\\bar{s}_0$, but the width of the distribution is wider, and hence discrimination between models is more diffcult, for lower $\\bar{s}_0$ (see fig.~\\ref{fig:sens}c and table~\\ref{table:opt}). Thus, in order to determine the probability of obtaining a certain value of $X_C$ in a given model (I, UB, B), one must make an assumption regarding the value of $\\bar{s}_0$. As mentioned in \\S~\\ref{sec:Method}, and discussed in detail in \\cite{Waxman:1996hp}, $\\bar{s}(z=0)$ may be constrained by the number of \"repeaters\", the number of sources producing multiple events. The last column of table~\\ref{table:opt} demonstrates that the presence or absence of repeaters may provide additional constraints on the source density (beyond $\\bar{s}_0\\gtrsim10^{-5}{\\rm Mpc}^{-3}$) once the number of events observed above 40~EeV exceeds several hundreds. Figure~\\ref{fig:sens}d demonstrates that the distribution of $X_{C}$ values, which is expected to be measured from the distribution of UHECR events with energies inferred {\\it by the experiment} to exceed a certain threshold, is sensitive to uncertainties in the experimental calibration of the absolute energy scale: Since the anisotropy is stronger at higher energy, a systematic underestimate (overestimate) of event energies leads to a shift of the $X_C$ distribution towards larger (smaller) values. Since the anisotropy is also stronger (at a given energy) for stronger bias, this implies that an underestimate (overestimate) of the absolute energy scale will lead to an overestimate (underestimate) of the bias of the UHECR source distribution. In order to distinguish between different bias models, the systematic uncertainty in the absolute energy calibration of the experiments should be reduced to well below the current $\\simeq25\\%$. We have shown, using the $X_C$ statistic, that the recently published $>5.7\\times10^{19}$~eV Auger data are consistent with a source distribution that traces LSS, with some preference to an UHECR source distribution that is biased with respect to the galaxy distribution, and inconsistent with isotropy at $\\simeq98\\%$ CL (see table~\\ref{table:Auger}). We have noted, however, that the optimization of the energy threshold made by the Auger collaboration analysis raises the concern that the significance with which isotropy is ruled out may be overestimated (see the end of \\S~\\ref{sec:Auger}). In order to confirm our detection of a correlation with LSS, one may search for the anisotropy signature in the Auger data using a lower energy threshold, $4\\times10^{19}$~eV, or repeat the analysis with an energy threshold of $5.7\\times10^{19}$~eV using a larger data set (which will be accumulated on a 1~yr time scale). According to table~\\ref{table:opt}, the detection of $\\approx300$ events above $4\\times10^{19}$~eV is required in order to enable one to identify the predicted anisotropy signature, and to discriminate between different bias models, with statistical significance exceeding 99\\% CL. The experimental exposure required to accumulate this number of events is uncertain, due to the uncertainty in the absolute energy calibration of the UHECR experiments, which implies an uncertainty in the absolute normalization of the UHECR flux (\\cite{Bahcall:2002wi}, see also fig.~\\ref{fig:spectrum}). Adopting, for example, the absolute energy scale of the HiRes experiment, the average number of events detected above 40~EeV is 300 per year for the Auger exposure of $7000$~km$^2$~sr, and about 50 events per year for the planned Telescope Array exposure of $1200$~km$^2$~sr \\cite{TA:2009}. These numbers are reduced by a factor of approximately 2 if the preliminary absolute energy scale of the Auger experiment is chosen (see fig.~\\ref{fig:spectrum}). The numbers given in table~\\ref{table:opt} imply also that the exposure accumulated within a few years of observation is unlikely to increase the significance of the detection to $>99.9\\%$ CL, unless the UHECR source density is comparable (or larger) than that of galaxies." }, "0801/0801.2383_arXiv.txt": { "abstract": "In this work we use a sample of 318 radio-quiet quasars (RQQ) to investigate the dependence of the ratio of optical/UV flux to X-ray flux, $\\alpha_{\\rm ox}$, and the X-ray photon index, $\\Gamma_X$, on black hole mass, UV luminosity relative to Eddington, and X-ray luminosity relative to Eddington. Our sample is drawn from the literature, with X-ray data from \\emph{ROSAT} and \\emph{Chandra}, and optical data mostly from the SDSS; 153 of these sources have estimates of $\\Gamma_X$ from \\emph{Chandra}. We estimate $M_{BH}$ using standard estimates derived from the H$\\beta$, Mg II, and C IV broad emission lines. Our sample spans a broad range in black hole mass ($10^6 \\lesssim M_{BH} / M_{\\odot} \\lesssim 10^{10}$), redshift ($0 < z < 4.8$), and luminosity ($10^{43} \\lesssim \\lambda L_{\\lambda} (2500$\\AA$) [{\\rm erg\\ s^{-1}}] \\lesssim 10^{48}$). We find that $\\alpha_{\\rm ox}$ increases with increasing $M_{BH}$ and $L_{UV} / L_{Edd}$, and decreases with increasing $L_X / L_{Edd}$. In addition, we confirm the correlation seen in previous studies between $\\Gamma_X$ and $M_{BH}$ and both $L_{UV} / L_{Edd}$ and $L_X / L_{Edd}$; however, we also find evidence that the dependence of $\\Gamma_X$ of these quantities is not monotonic, changing sign at $M_{BH} \\sim 3 \\times 10^8 M_{\\odot}$. We argue that the $\\alpha_{\\rm ox}$ correlations imply that the fraction of bolometric luminosity emitted by the accretion disk, as compared to the corona, increases with increasing accretion rate relative to Eddington, $\\dot{m}$. In addition, we argue that the $\\Gamma_X$ trends are caused by a dependence of X-ray spectral index on $\\dot{m}$. We discuss our results within the context of accretion models with comptonizing corona, and discuss the implications of the $\\alpha_{\\rm ox}$ correlations for quasar feedback. To date, this is the largest study of the dependence of RQQ X-ray parameters on black hole mass and related quantities, and the first to attempt to correct for the large statistical uncertainty in the broad line mass estimates. ", "introduction": "\\label{s-intro} The extraordinary activity associated with quasars involves accretion onto a supermassive black hole (SMBH), with the UV/optical emission arising from a geometrically thin, optically thick cold accretion disk \\citep{shakura73}, and the X-ray continuum arising from a hot, optically thin corona that Compton upscatters the disk UV photons \\citep[e.g.,][]{haardt91}. In highly accreting objects, like quasars \\citep[$0.01 \\lesssim L_{bol} / L_{Edd} \\lesssim 1$, e.g.,][]{woo02,vest04,mclure04,koll06}, the X-ray plasma geometry is expected to be that of a hot, possibly patchy, ionized `skin' that sandwiches the cold disk \\citep[e.g.,][]{bis77,liang77,nayak00}. However, the evidence for this is not conclusive, and relies on data from X-ray binaries and low-$z$ sources \\citep[e.g., see the dicussion by ][]{czerny03}. Other geometries are possible, including an accretion disk that evaporates into a hot inner flow \\citep[e.g.,][]{shap76,zdz99}, or a combination of a hot inner flow and a corona that sandwiches the disk \\citep[e.g.,][]{pout97,sob04a}. Furthermore, radiation pressure can drive an outflow from the disk into the corona if the two are cospatial, thus altering the physics of the corona \\citep{proga05}. Investigations of how quasar X-ray parameters depend on black hole mass, $M_{BH}$, and accretion rate relative to Eddington, $\\dot{m}$, offer important constraints on models of the disk/corona system. There have been attempts to link the evolution of SMBHs to analytic and semi-analytic models of structure formation \\citep[e.g.,][]{kauff00,hatz03,bromley04}, where black holes grow by accreting gas funneled towards the center during a galaxy merger until feedback energy from the SMBH expels gas and shuts off the accretion process \\citep[e.g.,][]{silk98,fabian99,wyithe03,begel05}. This `self-regulated' growth of black holes has recently been successfully applied in smoothed particle hydrodynamics simulations \\citep{dimatt05,spring05}. Within this framework, the AGN or quasar phase occurs during the episode of significant accretion that follows the galaxy merger, persisting until feedback from the black hole `blows' the gas away \\citep[e.g.,][]{hopkins06}. Hydrodynamic calculations have shown that line pressure is more efficient than thermal pressure at driving an outflow \\citep{proga07}, and therefore, the efficiency of AGN feedback depends on the fraction of energy emitted through the UV/disk component as compared to the X-ray/corona component. If the fraction of energy emitted in the UV as compared to the X-ray depends on $M_{BH}$ or $\\dot{m}$, then it follows that the efficiency of AGN feedback will also depend on $M_{BH}$ and $\\dot{m}$. This has important consequences for models of SMBH growth, as the SMBH may become more or less efficient at driving an outflow depending on its mass and accretion rate. Studies of the dependence of quasar X-ray/UV emission on black hole mass and accretion rate are therefore important as they allow us to constrain a $M_{BH}$- or $\\dot{m}$-dependent feedback efficiency. Numerous previous studies have searched for a luminosity and redshift dependence of $\\alpha_{ox} = -0.384 \\log L_X / L_{UV}$, the ratio of X-ray to UV/optical flux \\citep[e.g.,][]{avni82,wilkes94,yuan98,vig03b,strat05,steffen06,kelly07c}, and $\\Gamma_X$, the X-ray spectra slope \\citep[e.g.,][]{reeves00,bech03,dai04,ris05,grupe06}. Most studies have found a correlation between $\\alpha_{\\rm ox}$ and UV luminosity, $L_{UV}$, while the existence of a correlation between $\\alpha_{\\rm ox}$ and $z$ is still a matter of debate \\citep[e.g.,][]{bech03,vig03b,steffen06,just07,kelly07c}. In addition, studies of $\\Gamma_X$ have produced mixed results. Some authors have claimed a correlation between $\\Gamma_X$ and luminosity \\citep[e.g.,][]{bech03,dai04} or redshift \\citep[e.g.,][]{reeves97, vig99, page03}, while, others find no evidence for a correlation between $\\Gamma_X$ and $L_{UV}$ or $z$ \\citep[e.g.,][]{vig05,ris05,kelly07c}. A correlation between $\\Gamma_X$ and the $FWHM$ of the H$\\beta$ line has also been found \\citep[e.g.,][]{boller96,brandt97}, suggesting a correlation between $\\Gamma_X$ and black hole mass or Eddington ratio \\citep[e.g.,][]{laor97,brandt98}. Recently, it has become possible to obtain estimates of $M_{BH}$ for broad line AGN by calibrating results from reverberation mapping \\citep{peter04,kaspi05} for use on single-epoch spectra \\citep{wand99,vest02,mclure02,vest06,kelly07b}. This has enabled some authors to confirm a correlation between $\\Gamma_X$ and either $M_{BH}$ or $L_{bol} / L_{Edd}$ \\citep[e.g.,][]{lu99,gier04,porquet04,picon05,shemmer06}, where the X-ray continuum hardens with increasing $M_{BH}$ or softens with increasing $L_{bol} / L_{Edd}$. In addition, previous work has also found evidence for quasars becoming more X-ray quiet as $M_{BH}$ or $L_{bol} / L_{Edd}$ increase \\citep{brunner97,wang04}; however, studies involving the dependence of $\\alpha_{\\rm ox}$ on $M_{BH}$ or $L_{bol} / L_{Edd}$ have remained rare compared to studies of $\\Gamma_X$. It is important to note that the correlations inferred in previous work generally employ broad line mass estimates in combination with a constant bolometric correction. Therefore, most of the correlations found in previous work are, strictly speaking, between $\\Gamma_X$ or $\\alpha_{\\rm ox}$ and the estimates $M_{BH} \\propto L_{\\lambda}^{\\gamma} FWHM^2$ and $L_{bol} / L_{Edd} \\propto L_{\\lambda}^{1 - \\gamma} FWHM^{-2}$, where $\\gamma \\sim 0.5$. In this work, we investigate the dependence of $\\alpha_{\\rm ox}$ and $\\Gamma_X$ on black hole mass, optical/UV luminosity relative to Eddington, and X-ray luminosity relative to Eddington. We combine the main Sloan Digital Sky Survey (SDSS) sample of \\citet{strat05} with the sample of \\citet{kelly07c}, creating a sample of 318 radio-quiet quasars (RQQ) with X-ray data from \\emph{ROSAT} and \\emph{Chandra}, and optical spectra mostly from the SDSS; 153 of these sources have estimates of $\\Gamma_X$ from \\emph{Chandra}. Because the X-ray emission in radio-loud sources can have an additional component from the jet \\citep[e.g.,][]{zam81,wilkes87}, we focus our analysis on the radio-quiet majority. Our sample has a detection fraction of $87\\%$ and spans a broad range in black hole mass ($10^6 \\lesssim M_{BH} / M_{\\odot} \\lesssim 10^{10}$), redshift ($0 < z < 4.8$), and luminosity ($10^{43} \\lesssim \\lambda L_{\\lambda} (2500$\\AA$) [{\\rm erg\\ s^{-1}}] \\lesssim 10^{48}$), enabling us to effectively look for trends regarding $\\alpha_{\\rm ox}$ and $\\Gamma_X$. The outline of this paper is as follows. In \\S~\\ref{s-sample} we describe the construction of our sample, and in \\S~\\ref{s-specfits} we describe the procedure we used to fit the optical continuum and emission lines. In \\S~\\ref{s-mbhredd_est} we describe how we obtain broad line mass estimates, our bolometric correction, and argue that a constant bolometric correction provides a poor estimate of the bolometric luminosity. In \\S~\\ref{s-alfox_bh} we describe the results from a regression analysis of $\\alpha_{\\rm ox}$ on $M_{BH}, L_{UV} / L_{Edd},$ and $L_X / L_{Edd}$, and in \\S~\\ref{s-gamx_bh} we report evidence for a non-monotonic dependence of $\\Gamma_X$ on either $M_{BH}, L_{UV} / L_{Edd},$ and $L_X / L_{Edd}$. In \\S~\\ref{s-discussion} we discuss our results within the context of AGN disk/corona models, and we discuss the implications for a dependence of quasar feedback efficiency on black hole mass or accretion rate. In \\S~\\ref{s-summary} we summarize our main results. We adopt a cosmology based on the the WMAP best-fit parameters \\citep[$h=0.71, \\Omega_m=0.27, \\Omega_{\\Lambda}=0.73$,][]{wmap}. For ease of notation, we define $L_{UV} \\equiv \\nu L_{\\nu} (2500$\\AA$), L_X \\equiv \\nu L_{\\nu} (2\\ {\\rm keV}), l_{UV} \\equiv \\log \\nu L_{\\nu} (2500 $\\AA$),$ and $m_{BH} \\equiv \\log M_{BH} / M_{\\odot}$. ", "conclusions": "\\label{s-discussion} Previous work has found evidence for a correlation between $\\alpha_{\\rm ox}$ and both $M_{BH}$ \\citep{brunner97} and $L_{bol} / L_{Edd}$ \\citep{wang04}, and for a correlation between $\\Gamma_X$ and both $M_{BH}$ \\citep[e.g.,][]{porquet04,picon05} and $L_{bol} / L_{Edd}$ \\citep[e.g.,][]{lu99,gier04,shemmer06}, in aggreement with the results found in this work. However, our study differs from previous work in that we study a large sample of RQQs (318 sources with $\\alpha_{\\rm ox}$, 153 with $\\Gamma_X$) over a broad range in black hole mass ($10^6 \\lesssim M_{BH} / M_{\\odot} \\lesssim 10^{10}$) and redshift $0 < z < 4.8$; to date, this is the largest study of the dependence of the X-ray properties of RQQs on $M_{BH}, L_{UV} / L_{Edd},$ and $L_X / L_{Edd}$. In addition, this the first study of its kind to correct for the intrinsic statistical uncertainty in broad line mass estimates when quantifying the \\emph{intrinsic} trends between the X-ray emission and $M_{BH}, L_{UV} / L_{Edd},$ and $L_X / L_{Edd}$. Currently, there are two main types of geometries being considered for the comptonizing corona. The first of these is that of a `slab'-type geometry, possibly patchy, that sandwiches the disk \\citep[e.g.,][]{bis77,gal79,nayak00,sob04b}, and the second is that of a hot spherical inner advection dominated flow \\citep[e.g.,][]{shap76,zdz99}; hybrids between the two geometries have also been considered \\citep[e.g.,][]{pout97,sob04a}. There is a growing body of evidence that the advection dominated hot inner flow does not exist in objects with Eddington ratios $L_{bol} / L_{Edd} \\gtrsim 0.01$, as inferred from the existence of a relativistically broadened iron line \\citep[e.g.,][]{mineo00,lee02,fabian02}, relativistically broadened reflection of ionized material \\citep{janiuk01}, and by analogy with galactic black holes \\citep[e.g.,][]{esin97,nowak02}. The range in Eddington ratios probed by our study is at most $0.03 \\lesssim L_{bol} / L_{Edd} \\lesssim 2$, with a mean of $L_{bol} / L_{Edd} \\sim 0.25$. Therefore, the RQQs in our study are likely to have disks that extend approximately down to the last marginally stable orbit, and thus should only have the `slab' type geometries. \\subsection{Dependence of $\\alpha_{\\rm ox}$ on $M_{BH}$} \\label{s-alfox_mbh} In this work we have found that RQQs become more X-ray quiet as $M_{BH}$ increases, and confirmed the well-established relationship between $\\alpha_{\\rm ox}$ and $L_{UV}$. Because $L_{UV}$ increases with $M_{BH}$ and the accretion rate relative to Eddington, $\\dot{m}$, the well-known $\\alpha_{\\rm ox}$--$L_{UV}$ correlation is likely driven by the $\\alpha_{\\rm ox}$--$M_{BH}$ and $\\alpha_{\\rm ox}$--$\\dot{m}$ correlations. A correlation between $\\alpha_{\\rm ox}$ and $M_{BH}$ is expected even if the fraction of the bolometric luminosity emitted by the disk is independent of $M_{BH}$, as the effective temperature of the disk depends on $M_{BH}$. As $M_{BH}$ increases, the effective temperature of the disk decreases, thus shifting the peak of the disk emission toward longer wavelengths. Because the flux density at $2500$\\AA\\ lies redward of the peak in the disk SED over most of the range $M_{BH}$ probed by our study, this shift in the disk SED toward longer wavelengths produces an increase in $L_{UV}$ relative to $L_X$. We can use the standard thin disk solution to assess the evidence that the fraction of energy emitted by the corona depends on $M_{BH}$. We assume a simple model where the spectrum for the disk emission is that expected for an extended thin accretion disk, and the spectrum for the corona emission is a simple power-law with exponential cutoffs at the low and high energy end. According to \\citet{wandel00}, the spectrum from a radially extended thin accretion disk can be approximated as \\begin{equation} f^{D}_{\\nu} \\approx A_{D} \\left(\\frac{\\nu}{\\nu_{co}}\\right)^{-1/3} e^{-\\nu / \\nu_{co}}, \\label{eq-thindisk} \\end{equation} where $A_{D}$ is the normalization and $\\nu_{co}$ is the cut-off frequency. In this work, we choose the normalization to ensure that Equation (\\ref{eq-thindisk}) integrates to unity, and therefore $f^D_{\\nu}$ gives the shape of the disk emission. For a Kerr black hole, \\citet{malkan91} finds that the cut-off frequency is related to $M_{BH}$ as \\begin{equation} h \\nu_{co} = (6 {\\rm eV}) \\dot{m}^{1/4} (M_{BH} / 10^8 M_{\\odot})^{-1/4}. \\label{eq-cutoff} \\end{equation} We assume that the X-ray emission from the corona can be described by a simple power law with an exponential cutoff at the high and low end: \\begin{equation} f^{C}_{\\nu} = A_C \\nu^{-(\\Gamma_X - 1)} e^{-\\nu / \\nu_{high}} e^{-\\nu_{low} / \\nu}. \\label{eq-corona} \\end{equation} Here, $A_C$ is the corona spectrum normalization, $\\nu_{high}$ is the high energy cutoff, and $\\nu_{low}$ is the low energy cutoff. We choose the low energy cutoff to be $\\nu_{low} = 20\\ {\\rm eV}$, and we choose the high energy cutoff to be $\\nu_{high} = 200\\ {\\rm keV}$ \\citep[e.g.,][]{gilli07}. As with Equation (\\ref{eq-thindisk}), we choose the normalization in Equation (\\ref{eq-corona}) to be equal to unity. Denoting $f_D$ to be the fraction of bolometric luminosity emitted by the disk, our model RQQ spectrum is then \\begin{equation} L_{\\nu} \\approx L_{bol} \\left[f_D f^D_{\\nu} + (1 - f_D) f^C_{\\nu} \\right]. \\label{eq-modelspec} \\end{equation} We computed Equation (\\ref{eq-modelspec}) assuming a value of $\\dot{m} = 0.2$ and $f_D = 0.85$. We chose the value of the $\\dot{m} = 0.2$ because it is representative of the RQQs in our sample, and we chose the value $f_D = 0.85$ because it gives values of $\\alpha_{\\rm ox}$ typical of the RQQs in our sample. We vary $M_{BH}$ but keep $f_D$ and $\\dot{m}$ constant because we are interested in investigating whether there is evidence that assuming independence between $M_{BH}$ and both $f_D$ and $\\dot{m}$ is inconsistent with our $\\alpha_{\\rm ox}$ results. We compute Equation (\\ref{eq-modelspec}) for two forms of the dependence of $\\Gamma_X$ on $M_{BH}$. For the first model, we assume a constant value of $\\Gamma_X = 2$. For the second model, we assume that $\\Gamma_X$ depends on $M_{BH}$ according to our best fit regression results, where $\\Gamma_X$ depends on $M_{BH}$ according to Equation (\\ref{eq-gamxmbh_hbeta}) for $M_{BH} \\lesssim 3 \\times 10^8 M_{\\odot}$, and $\\Gamma_X$ depends on $M_{BH}$ according to Equation (\\ref{eq-gamxmbh_civ}) for $M_{BH} \\gtrsim 3 \\times 10^8 M_{\\odot}$. We ignore the intrinsic dispersion in $\\Gamma_X$. In Figure \\ref{f-modelspec} we show the spectra computed from Equation (\\ref{eq-modelspec}) for RQQs with $M_{BH} / M_{\\odot} = 10^7, 10^8, 10^9,$ and $10^{10}$. The dependence of the location of the peak in the disk emission on $M_{BH}$ is clearly illustrated. In Figure \\ref{f-alfox_model} we compare the $\\alpha_{\\rm ox}$--$M_{BH}$ regression results with the dependence of $\\alpha_{\\rm ox}$ on $M_{BH}$ expected from Equation (\\ref{eq-modelspec}) for both $\\Gamma_X$--$M_{BH}$ models. Under the thin disk approximation, a correlation is expected between $\\alpha_{\\rm ox}$ and $M_{BH}$, even if the fraction of bolometric luminosity emitted by the disk is independent of $M_{BH}$. However, our data are inconsistent with the assumption that $f_D$ and $M_{BH}$ are independent, given the thin disk approximation. Under the assumption that $f_D$ and $M_{BH}$ are independent, the $\\alpha_{\\rm ox}$--$M_{BH}$ correlation is too flat, and a increase in the fraction of bolometric luminosity emitted by the disk with increasing $M_{BH}$ is needed to match the steeper observed dependence of $\\alpha_{\\rm ox}$ on $M_{BH}$. Alternatively, if $f_D$ increases with increasing $\\dot{m}$, as we argue in \\S~\\ref{s-alfox_mdot}, then a steeper $\\alpha_{\\rm ox}$--$M_{BH}$ correlation also results if $M_{BH}$ and $\\dot{m}$ are correlated. In this case, if $f_D$ increases with increasing $\\dot{m}$, and if $M_{BH}$ increases with increasing $\\dot{m}$, then $f_D$ will also increase with increasing $M_{BH}$, thus producing a steeper observed $\\alpha_{\\rm ox}$--$M_{BH}$ correlation. To the extant that Equations (\\ref{eq-thindisk})--(\\ref{eq-modelspec}) accurately approximate the spectral shape of RQQs, our data imply that either the fraction of bolometric luminosity emitted by the disk increases with increasing $M_{BH}$, that $M_{BH}$ and $\\dot{m}$ are correlated, or both. Some theoretical models have suggested that the fraction of bolometric luminosity emitted by the disk should depend on $\\dot{m}$, but be relatively insensitive to $M_{BH}$ \\citep[e.g.,][]{czerny03,liu03}. Therefore, while a significant dependence of $f_D$ on $M_{BH}$ is not predicted by these disk/corona models, these models are still consistent with the interpretation that a $M_{BH}$--$\\dot{m}$ correlation is driving the steeper $\\alpha_{\\rm ox}$--$M_{BH}$ correlation. Unfortunately, without accurate estimates of $\\dot{m}$ we are unable to distinguish between these two possibilities. A shift in the peak of the disk SED with $M_{BH}$ may also explain the dependence of $\\alpha_{\\rm ox}$ on redshift observed by K07. K07 speculated that the observed hardening of $\\alpha_{\\rm ox}$ with increasing $z$ at a given $L_{UV}$ may be due to a correlation between $\\alpha_{\\rm ox}$ and $M_{BH}$, manifested through a $M_{BH}$--$z$ correlation. At a given $L_{UV}$, an increase in $M_{BH}$ will result in an increase in $L_X$ relative to $L_{UV}$, assuming that $\\dot{m}$ is not strongly correlated with $M_{BH}$. This is because an increase in $M_{BH}$ decreases the temperature of the disk, shifting the peak in the disk SED toward the red, and thus increasing the luminosity at 2500\\AA. However, since K07 investigated the dependence of $\\alpha_{\\rm ox}$ on $z$ at a given $L_{UV}$, the luminosity at 2500\\AA\\ is held constant. Therefore, the overall disk emission must decrease in order to keep the luminosity at 2500\\AA\\ constant despite the increase in $M_{BH}$. As a result, an increase in $M_{BH}$ at a given $L_{UV}$ will result in an increase in the X-ray luminosity relative to the luminosity at 2500\\AA. Because $M_{BH}$ and $z$ are correlated in our flux limited sample (e.g., see Figure \\ref{f-mbh_vs_z}), an increase in $z$ will probe RQQs with higher $M_{BH}$. As a result, RQQs will become more X-ray loud with increasing $z$, at a given 2500\\AA\\ luminosity. Consequently, deeper surveys that probe a greater range of $M_{BH}$ should not see as strong of a correlation between $M_{BH}$ and $z$, thereby reducing the magnitude of a $\\alpha_{\\rm ox}$--$z$ correlation. Indeed, investigations based on samples that span a greater range in luminosity do not find evidence for a correlation between $\\alpha_{\\rm ox}$ and $z$ \\citep[e.g.,][]{steffen06,just07}, qualitatively consistent with our interpretation of a $\\alpha_{\\rm ox}$--$z$ correlation. \\subsection{Dependence of $\\alpha_{\\rm ox}$ on $\\dot{m}$} \\label{s-alfox_mdot} We have found that $\\alpha_{\\rm ox}$ increases with increasing $L_{UV} / L_{Edd}$, and decreases with increasing $L_X / L_{Edd}$. The mere existence of these correlations is not particularly interesting, as we would expect that the ratio of optical/UV luminosity to X-ray luminosity would increase as the fraction of optical/UV luminosity relative to Eddington increases, and vice versa for an increase in $L_X / L_{Edd}$. However, the relative magnitude of these dependencies carries some information regarding the dependence of $\\alpha_{\\rm ox}$ on $\\dot{m}$. A correlation between $\\alpha_{\\rm ox}$ and $L_{UV} / L_{Edd}$ implies that $L_{UV} / L_X$ increases as the quantity $f_{UV} \\dot{m}$ increases, where $f_{UV}$ is the fraction of bolometric luminosity emitted at 2500\\AA. Likewise, an anti-correlation between $\\alpha_{\\rm ox}$ and $L_X / L_{Edd}$ implies that $L_{UV} / L_X$ decreases as the quantity $f_X \\dot{m}$ increases, where $f_X$ is the fraction of bolometric luminosity emitted at 2 keV. If the fraction of the bolometric luminosity emitted by the disk, as compared to the corona, increases with increasing $\\dot{m}$, then we would expect a strong increase in $L_{UV} / L_X$ with the product $f_{UV} \\dot{m}$, resulting from the dual dependency of $L_{UV} / L_X$ on $f_{UV}$ and $\\dot{m}$. Furthermore, because the fraction of bolometric luminosity emitted by the disk should decrease with increasing $f_X$, then, if the fraction of bolometric luminosity emitted by the disk increases with increasing $\\dot{m}$, we would expect a weaker dependence of $L_{UV} / L_X$ on the quantity $f_X \\dot{m}$. This is because an increase in $\\dot{m}$ causes an increase in the disk emission relative to the corona emission, which will then work against the decrease in disk emission relative to the corona that results from an increase in $f_X$. The end result is a weaker dependence of $L_{UV} / L_X$ on the product $f_X \\dot{m}$. Indeed, this is what we observe, where $L_{UV} / L_X \\propto (L_{UV} / L_{Edd})^{2.5}$ and $L_{UV} / L_X \\propto (L_X / L_{Edd})^{-1.5}$. Therefore, we conclude that the disk emission relative to the corona emission increases with increasing $\\dot{m}$. This is in agreement with some models of corona with a slab geometry \\citep[e.g.,][]{czerny97,janiuk00,merloni02,liu03}, where the $\\alpha_{\\rm ox}$--$\\dot{m}$ correlation arises due to a dependency of the size of the corona on $\\dot{m}$. Our result that $\\alpha_{\\rm ox}$ is correlated with $L_{UV} / L_{Edd}$ and anti-correlated with $L_X / L_{Edd}$ is inconsistent with a constant bolometric correction to both the optical/UV and X-ray luminosities. Instead, an increase in $L_{UV} / L_X$ with increasing $\\dot{m}$ implies that the bolometric correction depends on $\\dot{m}$. Because we conclude that the fraction of bolometric luminosity emitted by the disk increases with increasing $\\dot{m}$, this implies that the bolometric correction to the optical/UV luminosity decreases with increasing $\\dot{m}$, while the bolometric correction to the X-ray luminosity increases with increasing $\\dot{m}$. The direction of this trend is consistent with the results of \\citet{vasud07}, who find that the bolometric correction to the X-ray luminosity increases with increasing $L_{bol} / L_{Edd}$. Similarly, we have found evidence that the fraction of bolometric luminosity emitted by the disk depends on $M_{BH}$, therefore implying that the bolometric correction also depends on $M_{BH}$. Even if the fraction of bolometric luminosity emitted by the disk is independent of $M_{BH}$, the bolometric correction will still depend on $M_{BH}$ because the location of the peak in the disk emission will shift toward longer wavelengths as $M_{BH}$ increases. As $M_{BH}$ varies, the luminosity at 2500\\AA\\ probes a different region of the quasi-blackbody disk emission, thereby producing a dependence of bolometric correction on $M_{BH}$. \\subsection{Implications for Black Hole Feedback} A significant amount of recent work suggests that radiative and mechanical feedback energy from AGN plays an important part in galaxy and supermassive black hole coevolution \\citep[e.g.,][]{fabian99,wyithe03,hopkins05}. Within the context of these models, a nuclear inflow of gas, possibly the result of a galaxy merger, feeds the SMBH, thus igniting a quasar. The SMBH grows until feedback energy from the quasar is able to drive out the accreting gas, thus halting the accretion process. Hydrodynamic calculations of accretion flows have shown that the efficiency of the quasar in driving an outflow depends on the fraction of energy emitted through he UV/disk component as compared to the X-ray/corona component \\citep{proga07}. The disk component produces luminosity in the UV, which is responsible for driving an outflow via radiation pressure on lines, whereas the corona component produces luminosity in the X-rays, which is responsible for driving an outflow via thermal expansion. Calculations by \\citet{proga07} have shown that radiation driving produces an outflow that carries more mass and energy than thermal driving. If the efficiency of black hole feedback depends on the quasar SED, any dependence on $M_{BH}$ and $\\dot{m}$ of the fraction of AGN energy emitted in the UV as compared to the X-ray has important consequences for models of black hole growth. Because we have found evidence that the fraction of bolometric luminosity emitted by the disk increases with increasing $\\dot{m}$ and $M_{BH}$, this implies that black holes becomes more efficient at driving an outflow with increasing $\\dot{m}$ and $M_{BH}$. However, the $\\alpha_{\\rm ox}$--$M_{BH}$ correlation may be due to the combination of both a correlation between $M_{BH}$ and $\\dot{m}$, and a dependence of the location peak in the disk SED on $M_{BH}$. If the fraction of energy emitted by the disk only depends weakly on $M_{BH}$, as some theoretical models have suggested \\citep[e.g.,][]{czerny03,liu03}, the fraction of energy emitted in the UV will still decrease with increasing $M_{BH}$ becuase the peak of the disk emission will shift away from the UV. In this case, at a given $\\dot{m}$ we would expect that black holes will become less efficient at driving an outflow with increasing $M_{BH}$. \\subsection{Dependence of $\\Gamma_X$ on $M_{BH}$ and $\\dot{m}$} In this work we have also found evidence that $\\Gamma_X$ and $M_{BH}, L_{UV} / L_{Edd},$ and $L_X / L_{Edd}$ are not statistically independent. Moreover, the dependence of $\\Gamma_X$ on black hole mass or Eddington ratio appears to follow a non-monotonic form, although the $\\Gamma_X$--$M_{BH}$ trend is weak compared to the dependency of $\\Gamma_X$ on Eddington ratio. For the $\\Gamma_X$--$M_{BH}$ relationship, the X-ray continuum hardens with increasing black hole mass until $M_{BH} \\sim 3 \\times 10^8 M_{\\odot}$, after which the X-ray continuum softens with increasing black hole mass. The opposite is true for the $\\Gamma_X$--$L_{UV} / L_{Edd}$ and $\\Gamma_X$--$L_X / L_{Edd}$ trends, and further work is needed to confirm this result. Previous studies have not seen this non-monotonic trend because they have only employed the H$\\beta$ emission line, and therefore their samples have been dominated by low-$z$, low-$M_{BH}$ sources. \\subsubsection{Selection Effects} \\label{s-seleff} It is unlikely that the dependence of $\\Gamma_X$ on $M_{BH}, L_{UV} / L_{Edd},$ or $L_X / L_{Edd}$ is due to redshifting of the observable spectra range. If this were the case, then as $M_{BH}$ increases, so does $z$ due to selection effects, and thus we would observe a decrease in $\\Gamma_X$ as the `soft excess' shifts out of the observed 0.3--7 keV spectral range, while the compton reflection component shifts into the observed spectral range. However, there are lines of evidence that suggest that the $\\Gamma_X$ correlation are not due to redshifting of the observable spectral region, and that at least some of the observed dependency of $\\Gamma_X$ on $M_{BH}, L_{UV} / L_{Edd},$ and $L_X / L_{Edd}$ is real. First, in \\S~\\ref{s-ztest} we tested whether a sample of nine $z > 1$ test sources with $M_{BH} \\lesssim 3 \\times 10^8 M_{\\odot}$ were better described by a regression fit using the $z > 1.5, M_{BH} \\gtrsim 3 \\times 10^8 M_{\\odot}$ sources, or by a regression fit using the $z < 1, M_{BH} \\lesssim 3 \\times 10^8 M_{\\odot}$ sources. We found that the test sources were better fit using the regression of similar $M_{BH}$, and therefore that the difference in the $\\Gamma_X$--$M_{BH}$ correlations primarily depends on $M_{BH}$. Second, similar trends at low redshift between $\\Gamma_X$ and $M_{BH}$ or $L_{bol} / L_{Edd}$ have been seen in other studies that only analyze the hard X-ray spectral slope \\citep[typically 2--12 keV, e.g.,][]{picon05,shemmer06}, and thus these studies are not effected by the soft excess. Third, the compton reflection hump is unlikely to shift into the observabable spectral range until $z \\sim 1$. However, the contribution to the inferred $\\Gamma_X$ from compton reflection at $z \\gtrsim 1$ is likely weak, if not negligible, as our $z \\gtrsim 1$ sources have $M_{BH} \\gtrsim 10^8 M_{\\odot}$ and are highly luminous, and therefore are expected to only have weak reflection components \\citep{mineo00,ball01,bianchi07}. There are two scenarios in which the non-monotonic behavior of $\\Gamma_X$ with $M_{BH}$ or Eddington ratio may be artificially caused by selection. We will focus on the Eddington ratio dependency, as it is the strongest; however, our argument also applies to $M_{BH}$. First, the intrinsic dependency of $\\Gamma_X$ on Eddington ratio could be linear with increasing intrinsic scatter at high $L_{bol} / L_{Edd}$. Then, an inferred non-monotonic trend would occur if we were to systematically miss quasars with high $L_{bol} / L_{Edd}$ and steep X-ray spectra. Alternatively, there could be no intrinsic dependency of $\\Gamma_X$ on Eddington ratio. In this case, we would infer a non-monotonic trend if we were to systematically miss quasars with steep X-ray spectra at low and high $L_{bol} / L_{Edd}$, and quasars with flat X-ray spectra at moderate $L_{bol} / L_{Edd}$. We do not consider it likely that the observed non-monotonic dependence of $\\Gamma_X$ on Eddington ratio is due solely to selection effects. K07 describes the sample selection for sources with $\\Gamma_X$. With the exception of some of the $z > 4$ quasars, all sources from K07 were selected by cross-correlating the SDSS DR3 quasars with public \\emph{Chandra} observations. Almost all SDSS sources in K07 had serendipitious \\emph{Chandra} observations, and therefore were selected without regard to their X-ray properties. K07 estimated $\\Gamma_X$ for all sources that were detected in X-rays at the level of $3\\sigma$ or higher. Therefore, the only additional criterion beyond the SDSS selection imposed by K07 is the requirement that the source had to be detected in X-ray, which was fulfilled by $90\\%$ of the quasars; the undetected sources were slightly more likely to be found at lower redshift, probably due to the presence of unidentified BAL quasars. As a result, the K07 sample selection function is essentially equivalent to the SDSS quasar selection function. Because the SDSS selects quasars based on their optical colors, the most likely cause of selection effects is the optical color selection. There is evidence that $\\Gamma_X$ is correlated with the slope of the optical continuum, where the X-ray continuum flattens (hardens) as the optical continuum steepens (softens) \\citep[][K07]{gall05}. The SDSS selection probability is lower for red sources \\citep{rich06}, so we might expect to systematically miss sources with smaller $\\Gamma_X$. However, for the two scenarios described above, this is opposite the trend needed to explain the $\\Gamma_X$--$L_X / L_{Edd}$ relationship, where we need to at least systematically miss sources with larger $\\Gamma_X$. Furthermore, the drop in SDSS selection efficiency with optical spectral slope only occurs at $2 < z < 4$ \\citep{rich06}, thus we would expect a redshift dependence for this selection effect. As we have argued above, and in \\S~\\ref{s-ztest}, the non-monotonic trends for $\\Gamma_X$ cannot be completely explained as the result of different redshift ranges being probed. \\subsubsection{Implications for Accretion Physics} \\label{s-accphys} The dependence of $\\Gamma_X$ on $L_{UV} / L_{Edd}$ and $L_X / L_{Edd}$ is likely due to a dependence of $\\Gamma_X$ on $\\dot{m}$. If these $\\Gamma_X$ correlations were due to a dependence of $\\Gamma_X$ on $f_{UV}$ or $f_X$, then we would expect opposite trends for $L_{UV} / L_{Edd}$ and $L_X / L_{Edd}$, as $f_{UV}$ and $f_X$ should be anti-correlated. The fact that the regression results for the $\\Gamma_X$--$L_{UV} / L_{Edd}$ and $\\Gamma_X$--$L_X / L_{Edd}$ relationships are similar implies that $\\Gamma_X$ depends on $\\dot{m}$, and at most only weakly on $f_{UV}$ or $f_X$. A non-monotonic dependence of $\\Gamma_X$ on $\\dot{m}$ is predicted from the accreting corona model of \\citet{janiuk00}, as well as a non-monotonic dependence of $\\Gamma_X$ on the viscosity \\citep{bech03}. In addition, $\\Gamma_X$ is expected to steepen with increasing optical depth \\citep[e.g.,][]{haardt91,haardt93,czerny03}. One could then speculate that the dependence of $\\Gamma_X$ on $M_{BH}$ or $\\dot{m}$ is due to a non-monotonic dependence of the corona optical depth on $\\dot{m}$, which may indicate a change in the structure of the disk/corona system at $\\sim 3 \\times 10^8 M_{\\odot}$ or some critical $\\dot{m}$. Recent work also suggests a non-monotonic dependence of the optical/UV spectral slope, $\\alpha_{UV}$, on $\\dot{m}$ \\citep{bonning07,davis07}. From this work, it has been inferred that the optical/UV continuum becomes more red with increasing $\\dot{m}$ until $L_{bol} / L_{Edd} \\approx 0.3$, after which the optical/UV continuum becomes more blue with increasing $\\dot{m}$. Assuming the bolometric corrections described in \\S~\\ref{s-eddrat}, the turnover in the $\\Gamma_X$--$L_{bol} / L_{Edd}$ relationship also occurs at $L_{bol} / L_{Edd} \\approx 0.3$. \\citet{bonning07} suggested that the turnover in spectral slope at $L_{bol} / L_{Edd} \\sim 0.3$ may be due to a change in accretion disk structure, where the inner part of the accretion disk becomes thicker due to increased radiation pressure \\citep{abram88}. \\citet{bonning07} performed a simple approximation to this `slim disk' solution and found that it is able to produce a non-monotonic trend between optical color and Eddington ration. If the inner disk structure changes at high $\\dot{m}$, this change could alter the corona structure, producing the observed trend between $\\Gamma_X$ and Eddington ratio. Unfortunately, current models for corona geometry make a number of simplifying assumptions, and do not yet predict a specific relationship between $\\alpha_{\\rm ox}, \\Gamma_X, M_{BH},$ and $\\dot{m}$. Ideally, full magneto-hydrodynamic simulations \\citep[e.g.,][]{devill03,turner04,krolik05} that include accretion disk winds \\citep[e.g.,][]{murray95,proga04} should be used to interpret the results found in this work. However, MHD simulations have not advanced to the point where they predict the dependence of $\\alpha_{\\rm ox}$ and $\\Gamma_X$ on quasar fundamental parameters, but hopefully recent progress in analytical descriptions of the magneto-rotational instability \\citep{pessah06,pessah07} will help to overcome some of the computational difficulties and facilitate further advancement. In this work we have investigated the dependence of $\\alpha_{\\rm ox}$ and $\\Gamma_X$ on black hole mass and Eddington ratio using a sample of 318 radio-quiet quasars with X-ray data from \\emph{ROSAT} \\citep{strat05} and \\emph{Chandra} \\citep{kelly07c}, and optical data mostly from the SDSS; 153 of these sources have estimates of $\\Gamma_X$ from \\emph{Chandra}. Our sample spans a broad range in black hole mass ($10^6 \\lesssim M_{BH} / M_{\\odot} \\lesssim 10^{10}$), redshift ($0 < z < 4.8$), and luminosity ($10^{43} \\lesssim \\lambda L_{\\lambda} (2500$\\AA$) [{\\rm erg\\ s^{-1}}] \\lesssim 10^{48}$). To date, this is the largest study of the dependence of RQQ X-ray parameters on $M_{BH}, L_{UV} / L_{Edd},$ and $L_X / L_{Edd}$. Our main results are summarized as follows: \\begin{itemize} \\item We show that $\\alpha_{\\rm ox}$ is correlated with $L_{UV} / L_{Edd}$ and anti-correlated with $L_X / L_{Edd}$. This result is inconsistent with a constant bolometric correction being applicable to both the optical/UV luminosity and the X-ray luminosity. This result, when taken in combination with recent work by \\citet{vasud07}, implies that constant bolometric corrections can be considerably unreliable and lead to biased results. Instead, we argue that $L_{UV} / L_X$ increases with increasing $\\dot{m}$ and increasing $M_{BH}$, therefore implying that the bolometric correction depends on $\\dot{m}$ and $M_{BH}$. \\item We performed a linear regression of $\\alpha_{\\rm ox}$ on luminosity, black hole mass, $L_{UV} / L_{Edd}$, and $L_X / L_{Edd}$, and found significant evidence that $\\alpha_{\\rm ox}$ depends on all four quantities: $L_{UV} / L_{X} \\propto L_{UV}^{0.31 \\pm 0.03}, L_{UV} / L_{X} \\propto M_{BH}^{0.43 \\pm 0.06}, L_{UV} / L_{X} \\propto (L_{UV} / L_{Edd})^{2.57 \\pm 0.45},$ and $L_{UV} / L_X \\propto L_{X} / L_{Edd}^{-1.48 \\pm 0.14}$. The dependence of $\\alpha_{\\rm ox}$ on $L_{UV}$ may be due to the dual dependence of $\\alpha_{\\rm ox}$ on $M_{BH}$ and $\\dot{m}$. Because we have attempted to correct for the statistical uncertainties in $\\alpha_{\\rm ox}$ and the broad line estimates of $M_{BH}$, these results refer to the \\emph{intrinsic} relationships involving $\\alpha_{\\rm ox}$ and $M_{BH}$, and are not merely the relationships between $\\alpha_{\\rm ox}$ and the broad line mass estimates, $\\hat{M}_{BL} \\propto L^{\\gamma} FWHM^2$. \\item A correlation between $\\alpha_{\\rm ox}$ and $M_{BH}$ is expected from the fact that the peak in the disk emission will shift to longer wavelengths as $M_{BH}$ increases, even if the fraction of the bolometric luminosity emitted by the disk does not change with $M_{BH}$. Using a simple model for RQQ spectra, we argue that the observed $\\alpha_{\\rm ox}$--$M_{BH}$ correlation is steeper than that expected if both $\\dot{m}$ and the fraction of bolometric luminosity produced by the disk are independent of $M_{BH}$. The observed $\\alpha_{\\rm ox}$--$M_{BH}$ relationship therefore implies that either the fraction of bolometric luminosity emitted by the disk increases with increasing $M_{BH}$, that $M_{BH}$ is correlated with $\\dot{m}$, or both. \\item A correlation between $\\alpha_{\\rm ox}$ and $\\dot{m}$ is predicted from several models of `slab'-type corona. We argue that the weaker dependence of $\\alpha_{\\rm ox}$ on $L_X / L_{Edd}$ implies that $L_{UV} / L_X$ increases with increasing $\\dot{m}$. Considering that the efficiency of quasar feedback energy in driving an outflow may depend on the ratio of UV to X-ray luminosity, a correlation between $\\alpha_{\\rm ox}$ and both $M_{BH}$ and $\\dot{m}$ has important consequences for models of black hole growth. In particular, if supermassive black holes become more X-ray quiet at higher $\\dot{m}$, they will become more efficient at driving away their accreting gas, thus halting their growth. \\item Because of a possible nonlinear dependence of $\\Gamma_X$ on $M_{BH}, L_{UV} / L_{Edd}$, or $L_X / L_{Edd}$, we performed seperate regressions for the black hole mass estimates obtained from each emission line. We confirmed the significant dependence of $\\Gamma_X$ on $L_{UV} / L_{Edd}$ and $L_X / L_{Edd}$ seen in previous studies as inferred from the broad line mass estimates based on the H$\\beta$ line; however, we also find evidence that the $\\Gamma_X$ correlations change direction when including the C IV line. In particular, for the H$\\beta$ sample, the X-ray continuum hardens with increasing $M_{BH}$, while for the C IV sample, the X-ray continuum softens with increasing $M_{BH}$. Similar but opposite trends are seen with respect to $L_{UV} / L_{Edd}$ and $L_X / L_{Edd}$, and we conclude that these relationships can be interpreted as resulting from a correlation between $\\Gamma_X$ and $\\dot{m}$. Results obtained from the Mg II line were too uncertain to interpret. We analyzed two test samples to argue that this non-monotonic behavior is not due to the different redshifts probed by the two samples, or to problems with the estimates of $M_{BH}$ derived from the two lines; the different trends may be due to the difference in $M_{BH}$ probed by the two samples. A non-monotonic dependence of $\\Gamma_X$ on $M_{BH}$ and/or $\\dot{m}$ may imply a change in the disk/corona structure, although a non-monotonic dependence of $\\Gamma_X$ on $\\dot{m}$ and the viscosity is predicted by some models of `slab'-type coronal geometries. \\end{itemize}" }, "0801/0801.3450_arXiv.txt": { "abstract": "A simple model of quintessential inflation with the modified exponential potential $e^{-\\alpha \\phi} \\left[A+\\left(\\phi-\\phi_0\\right)^2\\right]$ is analyzed in the braneworld context. Considering reheating via instant preheating, it is shown that the evolution of the scalar field $\\phi$ from inflation to the present epoch is consistent with the observational constraints in a wide region of the parameter space. The model exhibits transient acceleration at late times for $0.96 \\lesssim A \\alpha^2 \\lesssim 1.26$ and $271 \\lesssim \\phi_0\\, \\alpha \\lesssim 273$, while permanent acceleration is obtained for $2.3\\times10^{-8} \\lesssim A \\alpha^2 \\lesssim 0.98$ and $255 \\lesssim \\phi_0\\, \\alpha \\lesssim 273$. The steep parameter $\\alpha$ is constrained to be in the range $5.3 \\lesssim \\alpha \\lesssim 10.8$. ", "introduction": "The recent measurements of the Wilkinson Microwave Anisotropy Probe (WMAP)~\\cite{Bennett:2003bz,Spergel:2003cb,Hinshaw:2006ia,Page:2006hz,Spergel:2006hy,Hinshaw:2008kr,Dunkley:2008ie,Komatsu:2008hk} on the cosmic microwave background (CMB) made it clear that the current state of the Universe is very close to a critical density and that the primordial density perturbations that seeded large-scale structure in the Universe are nearly scale-invariant and Gaussian, which is consistent with the inflationary paradigm. Inflation is often implemented in models with a single or multiple scalar fields~\\cite{Lyth:1998xn}, which undergo a slow-roll period allowing an early accelerated expansion of the Universe. Furthermore, the Universe seems to exhibit an interesting symmetry with regard to the accelerated expansion, namely, it underwent inflation at early epochs and is believed to be accelerating at present. The current acceleration of the Universe is supported by observations of high redshift type Ia supernovae~\\cite{Astier:2005qq,Riess:2006fw} and, more indirectly, by observations of the CMB and galaxy clustering~\\cite{Spergel:2006hy,Komatsu:2008hk,Seljak:2006bg}. Within the framework of general relativity, cosmic acceleration should be sourced by an energy-momentum tensor which has a large negative pressure (dark energy)~\\cite{Sahni:1999gb,Padmanabhan:2002ji,Ratra:1987rm,Wetterich:1987fm,Frieman:1995pm,Ferreira:1997au,Zlatev:1998tr,Brax:1999yv,Barreiro:1999zs,Bento:2001yv}. Therefore, in order to comply with the logical consistency and observations, the standard model should be valid somewhere between inflation at early epochs and quintessence at late times. It is then natural to ask whether one can build a model to join these two ends without disturbing the thermal history of the Universe. Attempts have been made to unify both these concepts using models with a single scalar field~\\cite{Peebles:1998qn,Copeland:2000hn,Huey:2001ae,Majumdar:2001mm,Nunes:2002wz,Sami:2004xk,Leith:2007bu,Neupane:2007mu}, i.e., in which a single scalar field plays the role of the inflaton and quintessence - the so-called quintessential inflation. On the other hand, in recent years there has been increasing interest in the cosmological implications of a certain class of braneworld scenarios where the Friedmann equation is modified at very high energies. In particular, in the Randall-Sundrum type II (RSII) model~\\cite{Randall:1999vf} the square of the Hubble parameter, $H^2$, acquires a term quadratic in the energy density, allowing slow-roll inflation to occur for potentials that would be too steep to support inflation in the standard Friedmann-Robertson-Walker (FRW) cosmology~\\cite{Cline:1999ts,Kaloper:1999sm,Shiromizu:1999wj,Binetruy:1999ut,Flanagan:1999cu,Maartens:1999hf,Maartens:2003tw,Bento:2006sr}. Indeed, in a cosmological scenario in which the metric projected onto the brane is a spatially flat FRW model, the Friedmann equation in four dimensions reads (after setting the 4D cosmological constant to zero and assuming that inflation rapidly makes any dark radiation term negligible)~\\cite{Cline:1999ts} \\begin{equation} H^2 = {1 \\over 3\\, M_4^2}\\, \\rho\\, \\left[1 + {\\rho \\over 2 \\lambda}\\right]~. \\label{eq:Friedmann} \\end{equation} Here $M_4$ is the 4D reduced Planck mass and $\\rho\\equiv \\rho_{\\phi} = \\frac{1}{2}{\\dot\\phi}^2 + V(\\phi)$ in a Universe dominated by a single minimally coupled homogeneous scalar field. The brane tension $\\lambda$ relates the 4D and 5D Planck masses through the relation \\begin{equation} \\lambda = \\frac{3}{32 \\pi^2} \\frac{M_5^6}{M_4^2}~, \\label{eq:tension} \\end{equation} where $M_5$ is the 5D Planck mass. We notice that Eq.~(\\ref{eq:Friedmann}) reduces to the usual Friedmann equation at sufficiently low energies, \\mbox{$\\rho \\ll \\lambda$}, while at very high energies $H\\propto\\rho$. In this scenario, all matter fields are confined to the brane and, hence, inflation is driven by a 4D scalar field trapped on the brane with the usual equation of motion \\begin{equation} {\\ddot \\phi} + 3H {\\dot \\phi} + V'(\\phi) = 0~, \\label{eq:kg} \\end{equation} where the prime denotes the derivative with respect to the scalar field $\\phi$. From Eqs.~(\\ref{eq:Friedmann}) and (\\ref{eq:kg}) it becomes clear that the presence of the additional term $\\sim \\rho^2/\\lambda$ increases the damping experienced by the scalar field as it rolls down its potential. It has been shown that, in the RSII braneworld context, quintessential inflation can occur for a sum of exponentials or cosh potentials~\\cite{Majumdar:2001mm,Nunes:2002wz,Sami:2004xk}. In this paper we show that the modified exponential potential (hereafter we adopt natural units, $M_4=1$, unless stated otherwise) \\begin{equation} V(\\phi)=e^{-\\alpha \\phi} \\left[A+\\left(\\phi-\\phi_0\\right)^2\\right] \\label{eq:albrecht} \\end{equation} also leads to a successful quintessential inflation model. In the context of quintessence, this potential was first analyzed by Albrecht and Skordis (AS)~\\cite{Albrecht:1999rm}. Afterward, it has been extensively studied in the literature~\\cite{Barrow:2000nc,Skordis:2000dz,Barreiro:2003ua,Blais:2004vt,Barnard:2007ta}. Regarding its motivation, it is worth noticing that exponential potentials naturally appear in 4D field theories coming from string/M-theory~\\cite{Gasperini:2001pc}, where the scalar field $\\phi$ is typically identified with the dilaton field. As far as the origin of the polynomial factor is concerned, it can be associated with a nontrivial K\\\"ahler term in an effective 4D supergravity theory~\\cite{Copeland:2000vh}. The scalar $\\phi$ could also be associated with a modulus (radion) field in curled extra dimensions~\\cite{Albrecht:2001xt}, which need not be universally coupled to matter/gauge fields and, therefore, is not subject to quantum corrections~\\cite{Doran:2002bc}. The tracking properties of the AS potential are similar to the pure exponential, namely, it allows sufficient radiation domination during big bang nucleosynthesis (BBN), followed by matter domination. Nevertheless, due to the presence of the polynomial factor, the field evolves to quintessence dominance near the present epoch. One should notice that, in order to the transition to happen near the present, the parameter $\\phi_0$ must be suitably chosen. In other words, this model does not explain the so-called coincidence problem. However, the model displays an interesting feature: it can lead to both permanent and transient acceleration regimes. Permanent acceleration occurs for $A\\alpha^2<1$, when the field is trapped in the local minimum of the potential. Transient vacuum domination arises in two ways~\\cite{Barrow:2000nc}: when $A\\alpha^2<1$ and the $\\phi$ field arrives at the minimum of the potential with enough kinetic energy to roll over the barrier and resumes descending the potential where $\\phi \\gg \\phi_0$, or for $A\\alpha^2>1$, when the potential loses its local minimum. In the models mentioned above, inflation takes place when the pure exponential potential dominates the potential. The exit from inflation takes place naturally when the slow-roll conditions are violated because the high-energy brane corrections become unimportant. Moreover, these models belong to the category of nonoscillating models in which the standard reheating mechanism does not work. In this case, one can employ an alternative mechanism of reheating via gravitational particle production~\\cite{Ford:1986sy,Spokoiny:1993kt,Grishchuk:1990bj}. However, this mechanism is faced with difficulties associated with excessive production of gravity waves. Indeed, the reheating mechanism based upon this process is extremely inefficient. The energy density of the produced radiation is typically one part in $10^{16}$~\\cite{Copeland:2000hn} to the scalar field energy density at the end of inflation. As a result, these models have a prolonged kinetic regime during which the amplitude of primordial gravity waves is enhanced and the nucleosynthesis constraints are violated~\\cite{Sahni:2001qp}. These problems can be circumvented if one invokes an alternative method of reheating, namely, the so-called instant preheating~\\cite{Felder:1998vq,Kofman:1994rk,Kofman:1997yn}. This mechanism is quite efficient and robust, and is well suited to nonoscillating models~\\cite{Felder:1999pv}. The larger reheating temperature in this model results in a smaller amplitude of relic gravity waves which is consistent with the BBN bounds~\\cite{Sami:2004xk}. ", "conclusions": "We have analyzed a simple model of quintessential inflation in the RSII braneworld context with a modified exponential potential. One of the attractive features of the model is that it can lead to transient acceleration at late times. This is particularly welcome in string theoretical formulations in order to avoid the difficulties arising in the $S$-matrix construction at the asymptotic future in a de-Sitter space~\\cite{Hellerman:2001yi,Fischler:2001yj,Witten:2001kn}. Assuming that the Universe was reheated via the instant preheating mechanism, we have shown that the evolution of the scalar field from inflation till the present epoch is consistent with the observations in a wide region of the parameter space. Requiring that the model meets various cosmological constraints at the different stages of the evolution, we were able to constrain tightly its parameters, as summarized in Eqs.~(\\ref{eq:boundtrans1})-(\\ref{eq:boundM5f}). In view of the very constrained bounds we obtained from the inflationary period, it is useful to consider how we could circumvent them in a simple and natural way. For instance, theoretical predictions for the inflationary observables may be modified by the presence of fields that are heavier than the Hubble rate during inflation. In this case, the coupling of the inflaton field to such heavy fields introduce corrections~\\cite{Bartolo:2007hx} which can be larger than the second-order contributions in the slow-roll parameters. Another way to change predictions for these observables is to consider the more general framework of Gauss-Bonnet gravity. In the presence of the Gauss-Bonnet term, the value of the spectral index is determined by the Gauss-Bonnet coupling parameter and the tension of the brane and is independent of the slope of the potential, thereby bringing the scenario in closer agreement with the most recent observations~\\cite{Lidsey:2003sj,Tsujikawa:2004dm}." }, "0801/0801.4046_arXiv.txt": { "abstract": "Observations of local X-ray absorbers, high-velocity clouds, and distant quasar absorption line systems suggest that a significant fraction of baryons may reside in multi-phase, low-density, extended, $\\sim 100$ kpc, gaseous halos around normal galaxies. We present a pair of high-resolution SPH (smoothed particle hydrodynamics) simulations that explore the nature of cool gas infall into galaxies, and the physical conditions necessary to support the type of gaseous halos that seem to be required by observations. The two simulations are identical other than their initial gas density distributions: one is initialized with a {\\em standard} hot gas halo that traces the cuspy profile of the dark matter, and the other is initialized with a {\\em cored} hot halo with a high central entropy, as might be expected in models with early pre-heating feedback. Galaxy formation proceeds in dramatically different fashions in these two cases. While the standard cuspy halo cools rapidly, primarily from the central region, the cored halo is quasi-stable for $\\sim 4$ Gyr and eventually cools via the fragmentation and infall of clouds from $\\sim 100$ kpc distances. After 10 Gyr of cooling, the standard halo's X-ray luminosity is $\\sim 100$ times current limits and the resultant disk galaxy is twice as massive as the Milky Way. In contrast, the cored halo has an X-ray luminosity that is in line with observations, an extended cloud population reminiscent of the high-velocity cloud population of the Milky Way, and a disk galaxy with half the mass and $\\sim 50\\%$ more specific angular momentum than the disk formed in the low-entropy simulation. These results suggest that the distribution and character of halo gas provides an important testing ground for galaxy formation models and may be used to constrain the physics of galaxy formation. ", "introduction": "It is well known that in the absence of feedback the majority of baryons in galaxy-size dark matter halos ($M_{\\rm v} \\sim 10^{12}$ M$_\\odot$) should have cooled into halo centers over a Hubble time (e.g. White \\& Rees 1978; Katz et al. 1992; Benson et al. 2003). In contrast, only $\\sim 20 \\%$ of the associated baryons in Milky-Way size halos are observed to be in a cold, collapsed form (Maller \\& Bullock 2004; Mo et al. 2005; Fukugita \\& Peebles 2006; Nicastro et al. 2007). An understanding of the feedback processes that act to solve this galaxy ``overcooling'' problem is a major goal of galaxy formation today. It is not known if the unaccounted baryons exist primarily as plasmas within normal galaxy halos (Maller \\& Bullock 2004; Fukugita \\& Peebles 2006; Sommer-Larsen 2006) or if they have been largely expelled as a result of energetic blow-out (see, e.g., Oppenheimer \\& Dav{\\'e} 2006, for a discussion). Observations of local X-ray absorbers (Williams et al. 2005; Fang et al. 2006), high-velocity clouds (Collins et al. 2005; Thom et al. 2006; Peek et al. 2007), and distant quasar absorption line systems (Tumlinson \\& Fang 2005; Kacprzak et al. 2007; Tinker \\& Chen 2007), suggest that a significant fraction of the missing halo baryons may reside in multi-phase, extended, $\\sim 100$ kpc, gaseous halos of normal galaxies. However, any hot gas around disk galaxies must be relatively low density in order to evade X-ray emission bounds ($S_x < 10^{-14}$ erg cm$^{-2}$ s$^{-1}$ arcmin$^{-2}$; Rasmussen et al. 2006, Li et al. 2007). These results, together with the fairly high covering factors in cool clouds implied by absorption line studies ($\\sim 50 \\%$; Kacprzak et al. 2007), are suggestive of a model where normal galaxies are surrounded by extended, low-density hot ($\\sim 10^6$ K) halos that are filled with fragmented, pressure supported cool ($\\sim 10^4$ K) clouds (Maller \\& Bullock 2004). Independently, models aimed at explaining the optical properties of galaxies have relied increasingly on the idea that extended, quasi-stable hot gas halos develop around massive galaxies (Kere{\\v s} et~al. 2005; Bower et al. 2006, Croton et al. 2006). It is suggested that these hot halos may be quite susceptible to feedback mechanisms, which could stabilize the systems to cooling and help explain the observed bimodality in galaxy properties (Dekel \\& Birnboim 2006). Observational probes of the gaseous halos of galaxies provide a potential means of testing these ideas. Entropy injected from feedback mechanisms will alter the density distribution of halo gas and affect associated cooling rates (and thus X-ray emission) and the distribution of cooling clouds fragmenting within the hot halos. Similarly, early feedback or {\\em pre-heating} before the halo collapses can affect halo gas profiles in a related manner, with positive consequences for galaxy properties at $z=0$ (Mo \\& Mao 2002; Oh \\& Benson; Lu \\& Mo 2007). \\begin{figure*} \\center \\includegraphics[height=50mm]{kaufmann1.ps} \\includegraphics[height=50mm]{kaufmann2.ps} \\caption{Renderings of the total gas density after 10 Gyr of cooling in our simulations. {\\bf Left:} results for our \"standard\" halo, initialized with a low-entropy cuspy hot halo. {\\bf Right:} results for the \"cored\" halo, initialized with a high-entropy, low-density hot halo of the same mass. In the first case, gas is concentrated around the galaxy. In the second case, the dense gas is more extended and distinct clouds are visible out to $\\sim100$ kpc. The boxes are $400$ kpc on a side. \\label{pic} } \\end{figure*} ", "conclusions": "Testing the effects of different initial conditions can teach us about the types of gaseous halos that are required to match the observed distributions of halo gas around galaxies. These preliminary simulations show that a cored initial gas density profile with a high initial entropy (as expected in pre-heating scenarios) but without any other feedback can produce disk masses, cool gas distributions, and X-ray emission signals that are in better agreement with observations than a more standard cuspy halo case with low central entropy. These results suggest that the distribution and character of gaseous galactic halos provide a powerful tool for understanding the physics of galaxy formation." }, "0801/0801.3516_arXiv.txt": { "abstract": "At least one massive binary system containing an energetic pulsar, PSR B1259-63/SS2883, has been recently detected in the TeV $\\gamma$-rays by the HESS telescopes. These $\\gamma$-rays are likely produced by particles accelerated in the vicinity of the pulsar and/or at the pulsar wind shock, in comptonization of soft radiation from the massive star. However, the process of $\\gamma$-ray production in such systems can be quite complicated due to the anisotropy of the radiation field, complex structure of the pulsar wind termination shock and possible absorption of produced $\\gamma$-rays which might initiate leptonic cascades. In this paper we consider in detail all these effects. We calculate the $\\gamma$-ray light curves and spectra for different geometries of the binary system PSR B1259-63/SS2883 and compare them with the TeV $\\gamma$-ray observations. We conclude that the leptonic IC model, which takes into account the complex structure of the pulsar wind shock due to the aspherical wind of the massive star, can explain the details of the observed $\\gamma$-ray light curve. ", "introduction": "PSR B1259-63/SS 2883 is the best known massive binary system containing a classical radio pulsar in a highly eccentric orbit around the Be type star. The pulsar creates strong relativistic wind which prevents accretion from the wind of the massive star. So then, it is expected that high energy processes, which are characteristic for the isolated radio ($\\gamma$-ray) pulsars, can also occur in the inner pulsar magnetosphere of PSR B1259-63. Due to the short rotational period, equal to $47$ ms, and strong surface magnetic field, the pulsar produces a strong relativistic wind which, in contrary to isolated pulsars, interacts with the wind from the massive star. This interaction process is especially important at the periastron passage when the pulsar is immersed in the inhomogeneous massive star wind. As a result of this interaction process, a relativistic shock in the pulsar wind appears. Particles can be accelerated on the shock to $\\sim$ TeV energies. Therefore, it has been suspected, that the binary system PSR 1259-63/SS 2883 might produce observable high energy $\\gamma$-ray radiation. This expectation has been confirmed during the last periastron passage of the pulsar in 2004, when the HESS detectors discovered the TeV $\\gamma$-ray emission from this binary system (Aharonian et al.~2005). The production of X-ray and $\\gamma$-ray radiation in PSR B1259-63 system in the model of interaction of the pulsar and stellar winds was considered for the first time by Grove et al.~(1997), and Tavani et al.~(1997). In those works $e^\\pm$ pairs, escaping from the pulsar magnetosphere, are accelerated additionally at the pulsar wind termination shock and interact with the magnetic field and thermal radiation of the massive star. Moreover, the conditions for propagation and cooling of $e^+e^-$ pairs have been considered as a function of the orbital phase. The produced X-ray light curves were in good agreement with the observed X-ray flux variability. The photon spectra were fitted under the assumption that the massive star equatorial wind is inclined to the orbital plain with the inclination angle $i > 25^o$. According to these models, the Lorentz factors of $e^\\pm$ pairs in the pulsar wind should be of the order of $\\gamma_e \\sim 10^6$, while the wind magnetization parameter, which determines the magnetic field strength at the shock, should be $\\sigma \\sim 10^{-1} - 10^{-2}$ (Tavani et al.~1996). A similar model for this binary system has been also considered by Kirk, Ball \\& Skjaeraasen~(1999). These authors argued for the first time that also the TeV $\\gamma$-rays should be produced in PSR 1259-63/SS 2883 binary system in the IC scattering of soft radiation from the massive star. Models with domination of the adiabatic and radiative energy losses were considered. However, they predicted the maximum of the TeV emission around the periastron passage which seems to be inconsistent with the observations. However, very recently Khangulyan et al.~(2007) argued that under special conditions such a general leptonic model can explain the TeV $\\gamma$-ray light curve observed from this system. Also the hadronic model for production of high energy $\\gamma$-rays has been considered by Kawachi et al.~(2004) who postulate acceleration of hadrons at the termination shock in the pulsar wind. In this model the hadrons can pass to the dense equatorial wind of the massive star and produce pions decaying in $\\gamma$-rays. The $\\gamma$-ray light curve expected in this model shows two maxima corresponding to the situation when the pulsar crosses of the equatorial wind of the massive star and the local minimum at the periastron passage. Such a $\\gamma$-ray light curve describes much better the HESS observations. The hadronic model has been also recently discussed by Neronov \\& Chernyakova~(2007). The above mentioned leptonic models have not taken into account the influence of the effects connected with an aspherical wind of the Be star. Such stars are characterized by dense but slow equatorial winds and low density but fast poloidal winds. The pulsar orbiting around the massive star has to pass trough these different winds. When it is immersed in the equatorial wind, the shock appears relatively close to the neutron star surface. In contrast, in the polar wind, the shock extends to relatively close to the massive star surface. In this last case, particles accelerated at the shock propagate through a much stronger radiation field and the ICS process should occur more efficiently. Moreover, the $\\gamma$-rays produced close to the stellar surface can also be partially absorbed, initiating in this way the IC $e^{\\pm}$ pair cascade. The importance of the $\\gamma$-ray production in the cascades occurring close to the surface of the massive star has been considered in the past, but only in the case of the isotropic winds (e.g. Sierpowska \\& Bednarek 2005a). In this paper, we consider those leptonic processes in a much more complicated geometry, i.e. by applying the aspherical wind properties. We calculate the $\\gamma$-ray spectra and light curves expected in such a more complicated and realistic model. The results of our calculations are compared with the recent observations of this binary system by the HESS telescopes in the TeV $\\gamma$-ray energies. The preliminary results has been published in Sierpowska \\& Bednarek~(2005b) and more details of the calculations are included in the PhD thesis of Sierpowska-Bartosik (Sierpowska-Bartosik~2006). ", "conclusions": "We considered the model for production of high energy gamma rays in the binary system consisting of a Be type massive star and the energetic pulsar applying the details of complicated geometry of the massive star wind. The model predictions are compared with recent observations of the binary system PSR B1259-63/SS2883. In this system, due to the interaction of the pulsar wind and the massive star wind, a shock wave is created. The geometry of the shock is complex and changes with the orbital phase as the pulsar passes through equatorial or polar wind of the massive star. Such structure of the massive star wind is characteristic for a Be type star. In the considered system, the equatorial wind plane is inclined to the orbital one. $\\gamma$-rays are produced by leptons which are accelerated at the shock region and after that scatter thermal photons from the massive companion. Since the soft photon field is anisotropic, we take into account the details of the geometry of the binary system. The change of the separation between the stars with the phase of the binary and the complex structure of the massive star wind have also big impact on the conditions for acceleration of leptons and their interactions with the stellar radiation. The maximum energies of leptons and their energy losses are determined by the magnetic field strength at the shock region which in turn is defined by the magnetization parameter ($\\sigma$) of the pulsar wind. In our calculations a value of $\\sigma = 10^{-2}$ and $10^{-4}$ is considered. The presented model takes into account the complicated geometry of the system and still it has to depend on many parameters which describe the position of the orbital plane and the equatorial plane of the stellar wind with respect to the observer. In such a complicated geometrically model, leptons can be accelerated in some cases relatively close to the stellar surface. Therefore, the possible absorption of the $\\gamma$-rays produced in the IC scattering process has to be also taken into account. The optical depths for leptons on ICS within the pulsar wind region are clearly too low for effective production of $\\gamma$-ray photons. However, leptons, which arrive from the pulsar to the shock region, are trapped by turbulent magnetic field after the shock and can be additionally accelerated. If the shock region is relatively close to the massive star, IC scattering becomes important and high energy $\\gamma$-rays are produced effectively. We calculate the $\\gamma$-ray spectra produced in the shock region for power law spectra of leptons in two different geometrical models. By applying the sets of parameters, I: $\\Delta\\omega_d=5^o, \\, \\theta_d=20^o, \\, i_d=40^o$ and II: $\\Delta\\omega_d=20^o, \\, \\theta_d=20^o, \\, i_d=30^o$, we reach the general consistency with the TeV $\\gamma$-ray observations. The best fit of the gamma light curve measured by the HESS telescopes (Aharonian, et al.~2005) is obtained by applying the model I with $\\sigma=10^{-4}$. However, this fit has been reached assuming that the periastron position of the pulsar is defined by the angle $\\omega_{\\rm p} = 148^o$ contrary to the present estimates, i.e. $\\omega_{\\rm p}\\sim 138^o$. Note, however that the estimation of the angle $\\omega_{\\rm p}$ depends on massive star wind parameters such as the inclination of the disc (applied from Wex et al. 1998) and the mass loss rate, which in case of PSR B1259-63 was fixed to $5\\times 10^{-8}$ M$_\\odot$yr$^{-1}$, and which are in fact not very well known. The calculated $\\gamma$-ray spectra are in agreement with the TeV HESS data provided that the spectral index of the spectrum of leptons is $\\alpha_i = -2.5$. The slope and also the fluxes of the photon spectra do not change with orbital phase, as was reported by the HESS collaboration. The final light curve for model I ($\\sigma=10^{-4}, \\omega = 148^o$) is characterized by a maximum for the phases when the pulsar crosses the equatorial disc after periastron. This fit differs from the observational data for phases when pulsar approaches the equatorial disc. This can be due to the complex geometry and structure of the shock region when the pulsar wind interacts with the strong equatorial wind. The model assumes that the acceleration of leptons occurs locally at the shock. In reality, this acceleration process of leptons arriving to the specific place at the shock can extend to other parts of the shock. Such a non-local acceleration process is very difficult to consider, but it might be important and influence some details of the calculated $\\gamma$-ray light curves. Based on the reported calculations, we conclude that the general shapes of the TeV and GeV $\\gamma$-ray light curves close to the periastron should be quite similar. However, farther out from the periastron, the GeV emission should continue with larger flux. This feature can be tested by the future observations in the GeV energy range by the AGILE and GLAST telescopes." }, "0801/0801.1039_arXiv.txt": { "abstract": "In this paper we review the current predictions of numerical simulations for the origin and observability of the warm hot intergalactic medium (WHIM), the diffuse gas that contains up to 50 per cent of the baryons at $z\\sim 0$. During structure formation, gravitational accretion shocks emerging from collapsing regions gradually heat the intergalactic medium (IGM) to temperatures in the range $T\\sim 10^5 - 10^7$ K. The WHIM is predicted to radiate most of its energy in the ultraviolet (UV) and X-ray bands and to contribute a significant fraction of the soft X-ray background emission. While \\ion{O}{vi} and \\ion{C}{iv} absorption systems arising in the cooler fraction of the WHIM with $T\\sim 10^5 - 10^{5.5}$ K are seen in \\fuse\\ and \\hst\\ observations, models agree that current X-ray telescopes such as \\chandra\\ and \\xmm\\ do not have enough sensitivity to detect the hotter WHIM. However, future missions such as \\constellation\\ and \\xeus\\ might be able to detect both emission lines and absorption systems from highly ionised atoms such as \\ion{O}{vii}, \\ion{O}{viii} and \\ion{Fe}{xvii}. ", "introduction": "\\label{Introduction} The cosmic baryon abundance was first inferred by applying the theory of primordial nucleosynthesis to observations of light element abundances (e.g.\\ \\citep{olive2000}) and has been confirmed by observations of the cosmic microwave background radiation (CMB). A recent measurement from the Wilkinson Microwave Anisotropy Probe reveals that $\\Omega_{\\mathrm b} h^2 = 0.0223_{-0.0009}^{+0.0007}$, where the Hubble parameter is $h=0.73_{-0.04}^{+0.03}$ \\citep{spergel2007}. At $z> 2$ most of the baryons are thought to reside in the diffuse, photoionised intergalactic medium (IGM) with $T\\sim 10^4 - 10^5$ K, traced by the \\lya\\ forest\\footnote{Since the diffuse intergalactic medium is highly ionised, the baryon density inferred from \\lya\\ forest observations is inversely proportional to the assumed ionisation rate. Since independent constraints on the UV intensity are only accurate to a factor of few at best, we cannot prove that most of the baryons reside in the forest.} (e.g.\\ \\citealt{rauch1997}; \\citealt{weinberg1997}; \\citealt{schaye2001}). The gas traced by the \\lya\\ forest, together with the mass in galaxies, fully accounts for the cosmic baryon abundance at $z\\sim 2$ and confirms the prediction of nucleosynthesis theories and the CMB measurements. However, when the mass of stars, interstellar gas and plasma in clusters of galaxies at redshift $z\\sim 0$ is considered, only a small fraction of the mass budget can be accounted for (e.g.\\ \\citealt{persic1992}; \\citealt{fukugita1998}; \\citealt{fukugita2004}). While the gas traced by \\lya\\ forest may account for about a third of the low-$z$ baryons (e.g.\\ \\citealt{danforth2007}), the majority of the baryons remain invisible. Unless the baryon budget at high redshift has been overestimated by two different measurements, a large amount of the low redshift baryons must be ``missing''. One of the basic predictions of cosmological, gas-dynamical simulations is the distribution of baryons in the universe. How many baryons are locked in stars? How many reside in clusters and filaments? What is the state of the diffuse gas? Simulations of structure formation predict that the missing baryons problem is measure of our technological limitations rather than a real problem. In other words, the baryons are out there, but we cannot see them because they reside in gas that is too dilute to be detected in emission and too hot to be traced by \\lya\\ absorption. The idea of gravitational heating of the intergalactic gas was first suggested by \\citet{sunyaev1972} and subsequently developed (e.g.\\ \\citealt{Nath2001}; \\citealt{furlanettoloeb2004}; \\citealt{rasera2006}). While the largest structures in the universe form, the IGM is heated by gravitational shocks that efficiently propagate from the collapsing regions to the surrounding medium. Simulations predict that gas compressed and heated by shocks can reach temperatures of $10^8$ K in rich clusters of galaxies, while filaments and mildly overdense regions are heated to temperatures in the range $10^5$ to $10^7$ K (e.g.\\ \\citealt{Cen1999}; \\citealt{dave2001}; \\citealt{Cen2006}). The latter are commonly known as the Warm-Hot Intergalactic Medium (WHIM), which represents about half the total baryonic mass in the universe at $z\\sim 0$. At $T\\sim 10^5 - 10^7$ K, the IGM is collisionally ionised and becomes transparent to \\lya\\ radiation. As a consequence, the \\ion{H}{i} \\lya\\ forest does not trace the bulk of the gas mass at low redshift. Although the shock-heated IGM emits radiation in the UV and in the soft X-ray bands, the total energy radiated away by mildly overdense gas is orders of magnitude too small to be detected by current instruments. Similarly, the absorption from such a gas along the line of sight to a bright X-ray source is too weak to be resolved by current spectrographs (\\citealt{richter2008} - Chapter 3, this volume). In this Paper, we review the predictions of numerical simulations for the properties and the observability of the WHIM. In Sect.~\\ref{origin} we briefly mention the numerical techniques used to simulate the IGM and we discuss the mechanisms that produce the WHIM, namely heating by gravitational accretion shocks and other, non-gravitational, heating processes. A more detailed description of the numerical techniques themselves can be found in \\citet{dolag2008} - Chapter 12, this volume. Sect.~\\ref{background} describes the contribution of the WHIM emission to the soft X-ray background. The current predictions for the detectability of the WHIM in emission and absorption are discussed in Sect.~\\ref{emission} and in Sect.~\\ref{absorption} respectively. Sect.~\\ref{sz} briefly reports on the effect of the WHIM on the CMB radiation and Sect.~\\ref{bfield} discusses the magnetisation of the IGM. We draw our conclusions in Sect.~\\ref{conclusion}. ", "conclusions": "\\label{conclusion} The search for the ``missing baryons'' has opened the way to our understanding of the most elusive baryonic component in the universe: the hot diffuse gas that traces the cosmic web. Given the paucity of observational findings, most of what we know about the WHIM is based on numerical simulations. These predict that at $z<1$ up to 50 per cent of the IGM has been heated to temperatures in the range $T \\sim 10^5 - 10^7$ K. While most of the heating is provided by gravitational accretion shocks propagating out of regions undergoing gravitational collapse, non-gravitational processes such as galactic winds and AGN feedback may play an important role in high density regions, mostly by preventing cooling and regulating the intensity of the X-ray background. The observability of the WHIM by future facilities has received a lot of attention and a wealth of predictions have been made based on numerical results. The detection of the WHIM seems to be extremely challenging, if not impossible, with current instruments and a new generation of X-ray telescopes with higher sensitivity and spectral resolution, such as \\constellation\\ and \\xeus, is needed to study the WHIM in absorption. In addition to high sensitivity and spectral resolution, high spatial resolution on a large field of view, as proposed for the Explorer of Diffuse emission and Gamma-ray burst Explosions (\\edge), would be ideal for mapping the WHIM emission on the sky. Although most theoretical predictions claim optimistically that we will be able to detect the WHIM in absorption with \\constellation\\ and \\xeus, and maybe in emission with some luck, there is a long way to go before observations can confirm the simulation predictions, or show us a different picture altogether of what and where the WHIM is." }, "0801/0801.2916_arXiv.txt": { "abstract": "{This paper presents a method which can be used to calculate models of the global solar corona from observational data.} % {We present an optimization method for computing nonlinear magnetohydrostatic equilibria in spherical geometry with the aim to obtain self-consistent solutions for the coronal magnetic field, the coronal plasma density and plasma pressure using observational data as input.} { Our code for the self-consistent computation of the coronal magnetic fields and the coronal plasma solves the non-force-free magnetohydrostatic equilibria using an optimization method. Previous versions of the code have been used to compute non-linear force-free coronal magnetic fields from photospheric measurements in Cartesian and spherical geometry, and magnetostatic-equilibria in Cartesian geometry. We test our code with the help of a known analytic 3D equilibrium solution of the magnetohydrostatic equations. The detailed comparison between the numerical calculations and the exact equilibrium solutions is made by using magnetic field line plots, plots of density and pressure and some of the usual quantitative numerical comparison measures.} { {We find that the method reconstructs the equilibrium accurately, with residual forces of the order of the discretisation error of the analytic solution. The correlation with the reference solution is better than $99.9\\%$ and the magnetic energy is computed accurately with an error of $< 0.1 \\%$. } } { We applied the method so far to an analytic test case. We are planning to use this method with real observational data as input as soon as possible. } ", "introduction": "In the recent past, numerical methods based on optimization principles have been used for a number of problems associated with the calculation of solar MHD equilibria. \\citet{wheatland:etal00} were the first to suggest the use of an optimization method to calculate nonlinear force-free fields in the corona from photospheric measurements. Since then the optimization method has been extended in various ways, for example by improving certain aspects of the original method for force-free fields \\citep[e.g.][]{wiegelmann04}, by introducing additional features such as plasma pressure into the method \\citep[e.g.][]{wiegelmann:etal06b} or by reformulating the method in other geometries \\citep[e.g.][]{wiegelmann07}. Optimization methods have the advantage of being conceptually straightforward and are reasonably easy to implement \\citep[see e.g.]{wheatland:etal00,wiegelmann:etal03,inhester:etal06}. At the moment they also seem to be very competitive in terms of computational efficiency \\citep[e.g][]{schrijver:etal06,metcalf:etal07}. A slight disadvantage compared to, for example, the Grad-Rubin method \\citep[e.g.][]{amari:etal97,wheatland06,inhester:etal06,amari:etal06} is that they still lack the same degree of rigorous mathematical basis existing for other methods. In the present paper we describe a further extension of the optimization method to calculate magnetohydrostatic (MHS) equilibria (including pressure and gravity) in spherical geometry. This is important for calculating global models of the corona including information going beyond just the structure of the magnetic field. In section \\ref{equations} we describe the basic equations of the optimization method for problem in hand. We then give a brief description of the analytical 3D MHS equilibria that we use to test the numerical code in section \\ref{3DMHS} and present the test results in section \\ref{results}. Our conclusions are presented in section \\ref{conclusions}. \\section[]{Basic equations} \\label{equations} The magnetohydrostatic (MHS) equations are given by \\begin{eqnarray} (\\nabla \\times {\\bf B}) \\times {\\bf B} -\\mu_0\\nabla p -\\mu_0 \\, \\rho \\, \\nabla \\Psi &=& {\\mathbf{0}} \\label{1} \\\\ \\nabla \\cdot {\\bf B} &=& 0, \\label{2} \\end{eqnarray} where ${\\bf B}$ is the magnetic field, $p$ the plasma pressure, $\\rho$ the mass density and $\\Psi=-\\frac{G M_s}{r}$ the gravitational potential with the gravitational constant $G$, the solar mass $M_s$ and the distance from the sun's center $r$. We do not assume an equation of state for the coronal plasma, but leave $p$ and $\\rho$ to be independent quantities. To find a magnetic field $\\mathbf{B}$, plasma pressure $p$ and plasma density $\\rho$ satisfying Eqs. (\\ref{1}) and (\\ref{2}), we follow the spirit of the previous optimization methods \\citep[e.g.][]{wheatland:etal00,wiegelmann:etal03a,wiegelmann04,wiegelmann:etal06b} and define the functional \\begin{eqnarray} L({\\bf B},p,\\rho) & = & \\int \\Bigg[\\frac{\\left|(\\nabla \\times {\\bf B}) \\times {\\bf B} -\\mu_0 \\, \\nabla p -\\mu_0 \\, \\rho \\,\\nabla \\Psi \\right|^2}{B^2} \\nonumber \\\\ && \\; + \\; |\\nabla \\cdot {\\bf B}|^2\\Bigg] \\; \\, r^2 \\, \\sin \\theta \\, dr \\, d \\theta \\, d \\phi. \\label{defL_Bprho} \\end{eqnarray} It is obvious that Eqs. (\\ref{1}) and (\\ref{2}) are satisfied if $L=0$. Here ${\\bf B}$ is a vector field, but not necessarily a solenoidal magnetic field during the iteration. The numerical method is based on an iterative scheme to minimize the functional $L$. To simplify the mathematical derivation we define the quantities \\begin{eqnarray} {\\bf \\Omega_a} &=& \\; B^{-2} \\;\\left[(\\nabla \\times {\\bf B}) \\times {\\bf B} -\\mu_0 \\, \\nabla p -\\mu_0 \\,\\rho\\, \\nabla \\Psi \\right] \\\\ {\\bf \\Omega_b} &=& B^{-2} \\;\\left[(\\nabla \\cdot {\\bf B}) \\; {\\bf B} \\right], \\label{defomega} \\end{eqnarray} and rewrite $L$ as \\begin{equation} L=\\int_{V} \\; \\left[ B^2 \\Omega_a^2+ B^2 \\Omega_b^2 \\; r^2 \\right] \\, \\sin \\theta \\, dr \\, d \\theta \\, d \\phi. \\end{equation} Taking the derivative of $L$ with respect to an iteration parameter $t$, where ${\\bf B}, \\; p, \\; \\rho$ are assumed to depend upon $t$, we obtain \\begin{eqnarray} \\frac{1}{2} \\; \\frac{d L}{d t} &=&-\\int_{V} \\frac{\\partial {\\bf B}}{\\partial t} \\cdot {\\bf F} \\; dV +\\int_{V} \\frac{\\partial p}{\\partial t} \\; \\mu_0 \\nabla \\cdot {\\bf \\Omega_a} \\; dV \\nonumber \\\\ &&-\\int_{V} \\frac{\\partial \\rho}{\\partial t} \\; \\mu_0 {\\bf \\Omega_a} \\cdot \\nabla \\Psi \\; dV -\\int_{S} \\frac{\\partial {\\bf B}}{\\partial t} \\cdot {\\bf G} \\; dS \\nonumber \\\\ && -\\int_{S} \\frac{\\partial p}{\\partial t} \\; \\mu_0 {\\bf \\Omega_a} \\cdot {\\bf dS}, \\end{eqnarray} where \\begin{eqnarray} {\\bf F} & =& \\nabla \\times ({\\bf \\Omega_a} \\times {\\bf B} ) - {\\bf \\Omega_a} \\times (\\nabla \\times {\\bf B}) \\nonumber \\\\ && + \\nabla({\\bf \\Omega_b} \\cdot {\\bf B})- {\\bf \\Omega_b}(\\nabla \\cdot {\\bf B}) +(\\Omega_a^2 + \\Omega_b^2) \\, {\\bf B} \\end{eqnarray} and \\begin{eqnarray} {\\bf G} & = & {\\bf \\hat n} \\times ({\\bf \\Omega_a} \\times {\\bf B} ) -{\\bf \\hat n} (\\bf \\Omega_b \\cdot \\bf B). \\end{eqnarray} Assuming that \\begin{eqnarray} \\frac{\\partial {\\bf B}}{\\partial t} &=& \\mu \\; {\\bf F} \\label{Beq}\\\\ \\frac{\\partial p}{\\partial t} &=& -\\nu \\; \\mu_0 \\nabla \\cdot {\\bf \\Omega_a} \\label{peq} \\\\ \\frac{\\partial \\rho}{\\partial t} &=& \\xi \\; \\mu_0 \\; {\\bf \\Omega_a} \\cdot \\nabla \\Psi =\\xi \\; \\mu_0 \\frac{G M_s}{r^2} \\; {\\bf \\Omega_a} \\cdot {\\bf e_r} \\label{rhoeq} \\end{eqnarray} with positive definite parameters $\\mu$, $\\nu$ and $\\xi$ and that the boundary integrals vanish, one can easily see that $L$ is monotonically decreasing with $t$ (note that this does not necessarily imply that $L$ tends to zero). Discretized versions of Eqs. (\\ref{Beq}) to (\\ref{rhoeq}), together with appropriate boundary conditions, form the basis for the numerical scheme. The boundary conditions have to be consistent with the assumption that the boundary integrals vanish, for example, by keeping the magnetic field, pressure and density fixed on the boundaries during the iteration. For testing the method in this paper we shall take these boundary conditions from the known exact solution. For practical applications these would have to come from observational data. We remark that due to the introduction of additional forces the constraints on the consistency of the boundary conditions for the magnetic field are somewhat different from the force-free case. However, the general theory of magnetohydrostatic equilibria requires for example that the pressure has the same value at both foot points of a closed field line under the general conditions assumed in the present paper. This is similar to the Cartesian case discussed by \\citet{wiegelmann:etal06b}, where the pressure equation was forward integrated along field lines using an upwind method from one foot point to the other (in the test case described later we make use of the property that the pressure is known to be consistent). In the numerical implementation based on this method one has to choose the product of the time-step $\\Delta t$ with the three numerical parameters $\\mu$, $\\nu$ and $\\xi$. Usually these products have to be small enough to achieve convergence and this is ensured by an adaptive time-step control. In this paper we have chosen the three parameters (multiplied by $\\Delta t$) to have the same values on all grid points. Previous experience with applying a similar method to force-free magnetic fields in spherical geometry \\citep{wiegelmann07} showed that choosing the same values for the entire box can lead to long computing times in the polar regions due to the distortion of the numerical grid in spherical polar coordinates towards the poles (note that the poles $\\theta=0$ and $\\theta=\\pi$ are excluded from the computational domain). This could in principle be compensated by allowing for a spatial variation of the iteration parameters. ", "conclusions": "\\label{conclusions} \\begin{figure*} \\includegraphics[angle=-90, width=\\textwidth]{8244f4.eps} \\caption{Outlook: How can the tool become applied to data? The basic idea is to compute first a nonlinear force-free field from observed vector magnetograms and then model the plasma along the loops. Our newly developed magneto-hydrostatic code uses the resulting magnetic field and plasma configuration as input for the computation of a self-consistent equilibrium.} \\label{fig4} \\end{figure*} We have extended the optimization method originally proposed for the reconstruction of force-free magnetic fields \\citep{wheatland:etal00} to global magnetohydrostatic equilibria including the pressure force and the gravitational force in spherical geometry. The proposed generalization of the optimization method leads to two additional equations for the pressure and the density that have to be solved simultaneously to the magnetic field equation. Boundary conditions for the magnetic field, the pressure and the density are necessary to complete the problem. We have implemented a numerical code based on the proposed method and have tested the code using a known three-dimensional magnetohydrostatic equilibrium \\citep{neukirch95}. The numerical calculation is started from a potential field with the same radial magnetic field component as the analytical equilibrium on the boundary. The initial pressure and density distribution are a spherically symmetric stratified atmosphere in hydrostatic balance. Both visual inspection of the results as well as a quantitative analysis using various diagnostic measures indicate that the method works well and converges to the analytic equilibrium. For the presented tests we used a low spatial resolution and got a relatively long computing time (about 200 000 iteration steps) until convergence. In experiments with the force-free version of our spherical code \\citep[see][]{wiegelmann07} we found that the computing time scales with $N^{5.4}$ regarding the number of grid points $N$ in one spatial direction. This is somewhat slower as the theoretical estimate of $N^5$ for a cartesian optimization code obtained by \\citet{wheatland:etal00}. The spherical magnetohydrostatic code is significantly slower than the cartesian force-free code for two reasons. \\begin{enumerate} \\item The convergence of the numerical grid towards the poles requires a sufficiently small time-step. \\item The plasma $\\beta$ might vary strongly in the entire region and in particular low-$\\beta$-regions require very small time-steps to compute the magnetic field and plasma simultaneously, because small changes in the Lorentz force can result in considerably large changes in the low $\\beta$ plasma. \\end{enumerate} Point 1.) can be addressed by using a more sophisticated numerical grid, e.g. the Yin-Yang grid developed by \\citet{kageyama:etal04}, which has been applied in geophysics \\citep[see e.g.][]{yoshida:etal04}. This overset grid contains two complementary grids which lead to an almost uniformly spaced spherical grid. An additional advantage of the Yin-Yang grid is that it is suitable for massive parallelization. The Yin-Yang grid has been applied for geophysical simulations on the Earth simulator super-computer in Yokohama. To speed up the 2. point one might compute first the magnetic field alone as a nonlinear force-free field (which is a reasonable approximation in low-$\\beta$-regions) and switch on the self-consistent plasma iteration only after for a fine-tuning. A multi-scale approach, as recently implemented to speed up our force-free cartesian optimization code \\citep[see][for details]{metcalf:etal07} is also an option worth trying for the spherical implementation of our force-free and magnetohydrostatic codes. We are confident, that the above mentioned potential for improvements together with a massive parallelization will allow us to apply our newly developed method to real data with a reasonable grid resolution. In figure \\ref{fig4} we outline a scheme on how the code might be applied to data. {\\tw As boundary conditions on pressure and density are not directly measured, we propose the following approach.} The basic idea is to compute first a force-free magnetic field and then model the plasma along the magnetic loops, e.g., by the use of scaling laws and optionally with the help of a tomography code. Such an approach has been used by \\citet{schrijver:etal04} by using a global potential magnetic field and specifying free scaling law parameters (e.g., heating function) by comparing artificial plasma images (created by line-of-sight integration from the model plasma) with X-ray and EUV observations. We propose to generalize this approach by using a nonlinear force-free magnetic field model and compare the model plasma with observations from two viewpoints as provided by the STEREO-mission. Optional STEREO-images can be used additionally to approximate the coronal density distribution by a tomographic inversion. As a consequence of this step (modelling the plasma along a magnetic loop) the plasma pressure is consistent along the loops. Different values for the pressure on different field-lines will violate the force-free condition for the magnetic field, however, and the configuration is not exactly in a magnetohydrostatic equilibrium. Finite pressure gradients have to be compensated by a Lorentz force. This computation can be done with the help of the program described in this paper. We propose to use the force-free magnetic field configuration and the model plasma as initial state for our newly developed magnetohydrostatic code to compute a self-consistent MHS-equilibrium. For a low $\\beta$ plasma the back-reaction of the plasma onto the magnetic field will be small, for higher values of $\\beta$ (as found e.g., in helmet streamers) the magnetic field might change significantly. As a result of this approach one has reconstructed the 3D coronal magnetic field and plasma configuration self-consistently within the magnetohydrostatic approach and the model is consistent with measured photospheric vector magnetograms and observed coronal images from different viewpoints as well." }, "0801/0801.2257_arXiv.txt": { "abstract": "We analyze the mass distribution of cores formed in an isothermal, magnetized, turbulent, and self-gravitating nearly critical molecular cloud model. Cores are identified at two density threshold levels. Our main results are that the presence of self-gravity modifies the slopes of the core mass function (CMF) at the high mass end. At low thresholds, the slope is shallower than the one predicted by pure turbulent fragmentation. The shallowness of the slope is due to the effects of core coalescence and gas accretion. Most importantly, the slope of the CMF at the high mass end steepens when cores are selected at higher density thresholds, or alternatively, if the CMF is fitted with a lognormal function, the width of the log-normal distribution decreases with increasing threshold. This is due to the fact that gravity plays a more important role in denser structures selected at higher density threshold and leads to the conclusion that the role of gravity is essential in generating a CMF that bears more resemblance with the IMF when cores are selected with an increasing density threshold in the observations. ", "introduction": "It is now widely accepted that stars form in dense and gravitationally bound molecular cloud (MC) cores. Thus, understanding the properties of the core mass function (CMF) is the cornerstone of any successful theory of the stellar initial mass function (IMF). The question whether there is a 1-to-1 mapping between the CMF and the IMF remains an open one. The CMF is usually represented by $dN/dM=M^{\\alpha}$, where $\\alpha$ takes different values in different mass ranges (for the IMF, $\\alpha=-2.35$ for $M \\gtrsim 0.5$ M$_{\\odot}$, Salpeter 1955). The determination of the CMF of cores of several MC regions have been obtained using a variety of wavelengths and techniques such as mm spectroscopy of various molecular lines, submm/mm continuum emission from dust, and stellar near infrared absorption by dust. Usually, no prior assumptions are made on the type of structures identified in these studies. They can be a mixture of prestellar cores, unbound objects, and eventually protostellar envelopes. For nearby MCs, where it is possible to identify dense cores (with average number densities $n \\gtrsim 10^{4}$ cm$^{-3}$; hereafter Cores), the CMF is usually described by a two component power law above and below a turnover point (with exponents $\\alpha_{h}$ and $\\alpha_{l}$, respectively). Starting with the 1.2 mm study of the Ophiuchus starless dense cores by Motte et al. (1998), a number of submm/mm dust continuum observations have found $\\alpha_{h}$ and $\\alpha_{l}$ to be in the range [-2.1,-3] and [-1.3,-1.5], respectively (Motte et 2001; Testi \\& Sargent 2001; Johnstone et al. 2006; Beuther \\& Schilke 2004; Stanke et al. 2006; Reid \\& Wilson 2006; Johnstone \\& Bally 2006; Kirk et al. 2006; Enoch et al. 2006). These estimates bracket the values of the exponents of the IMF (Scalo 1998; Kroupa 2002; Chabrier 2003). Similar slopes are found using high density molecular tracers such as H$^{13}$CO$^{+}$ (Onishi et al. 2002; Ikeda et al. 2007) and the near infrared absorption technique (Alves et al. 2007). On the other hand, submm/mm observations of more distant clouds in which it is only possible to identify lower density structures ($n \\sim 10^{2}-10^{3}$ cm$^{-3}$; hereafter Clumps), exhibit values of $\\alpha_{h}$ in the range of [-1.4,-1.8] if the whole CMF is fitted with a single power law (Coppin et al. 2000; Kerton et al. 2001; Tothill et al. 2002; Mookerjea et al. 2004; Massi et al. 2007; Li et al. 2007, Mu\\~{n}oz et al. 2007) or slightly steeper ($\\alpha_{h} \\sim -2$) if it can be fitted by a two component power law (Reid \\& Wilson 2005). Similar slopes are obtained from molecular line observations using low/intermediate density molecular line tracers (Stutzki \\& G\\\"{u}tsen 1990; Lada et al. 1991; Blitz 1993; Brand \\& Wouterloot 1995; Kramer et al. 1998; Heithausen et al. 1998; Kawamura et al. 1998; Wilson et al. 2003). Values of $\\alpha_{h} \\sim [-1.6,-1.7]$ resemble the slopes of the stellar clusters mass function (Elmegreen 2000), which hints to the fact that these objects cannot be the direct progenitors of individual stars and must undergo additional fragmentation. The discrepancy in $\\alpha_{h}$ between the Cores and Clumps populations does not seem to be merely an effect of the spatial resolution and hence of mass resolution. Additional evidence might be that, in Motte et al. (1998), the inclusion of the larger size cores in the CMF changes the values of $\\alpha_{h}$ from $-2.5$ to $\\sim -1.5$. An explanation of the origin of the discrepancy would be that different physical processes dominate the evolution of these two populations of cores. The close resemblance between the Cores CMF and the IMF suggests that they are more likely to be gravitationally bound (which is confirmed by mm line observations, Andr\\'{e} et al. 2007) and are the direct progenitors of stars or small stellar systems, whereas Clumps are more likely to be dominated by turbulence. Note that the description of the CMF by power law functions is not the unique option. Reid el al. (2006) argued that a lognormal function is a better fit to the observed CMF in many regions and Goodwin et al. (2007) showed that the CMF of the submm cores obtained by Nutter \\& Ward-Thompson (2007) is well represented by a lognormal distribution (also Fig. 2 in Andr\\'{e} et al. 2008). Several physical processes are believed to play a role in shaping the CMF: gravitational and thermal fragmentation (Larson 1985; Klessen 2001; Li et al. 2003), supersonic turbulence (Padoan \\& Nordlund 2002, PN02), accretion (Larson 1992; Bonnell et al. 2001; Bate \\& Bonnell 2005), core coalescence (Murray \\& Lin 1996; Dib et al. 2007a,2008), feedback (Adams \\& Fatuzzo 1996), and magnetic fields (Shu et al. 2004). A number of numerical studies of MCs that include some of these physical processes have derived a CMF that bears some resemblance with the CMF of Cores and the IMF (Gammie et al. 2003; Li et al. 2004; Ballesteros-Paredes et al. 2006; Padoan et al. 2007) or with the CMF of Clumps (Hennebelle \\& Audit 2007, Heitsch et al. 2008). The origin of the differences between the Cores and Clumps CMFs and the eventual link between them has not been investigated so far. In this study, we identify cores in a high resolution 3D numerical simulation of a magnetized, turbulent, and self-gravitating MC model at different density thresholds. In Dib et al. (2007b), we have shown, using a detailed virial analysis that the inner parts of dense structures, identified at higher density thresholds, are more gravitationally bound than the same objects identified at lower thresholds. Thus, constructing the CMF at different thresholds is a direct test of the effects of gravity on the CMF. A striking example of how the density threshold can modify the inferred masses of cores is the case of the core L1512. Its derived mass from CO observations is $\\sim 6$ M$_{\\odot}$ (Falgarone et al. 2001), whereas it was estimated from submm observations at $\\sim 1.4$ M$_{\\odot}$ (Ward-Thompson et al. 1999). For the whole CMF, Kramer et al. (1998) derived a value of $\\alpha_{h}=-1.72 \\pm 0.09$ for the Orion B-S region from $^{13}$CO(1-0) observations whereas Johnstone et al. (2006) derived a value of $\\sim -2.5$ from submm observations. Another example is the case of the NGC 7538 region, observed by Kramer et al. (1998) to have a CMF slope of $\\alpha_{h}=-1.65 \\pm 0.05$ in the $^{13}$CO(1-0) line and $\\alpha_{h}=-1.79 \\pm 0.12$ in the C$^{18}$O(1-0) line which traces higher density material. ", "conclusions": "\\label{conc} In this work, we have constructed the mass function of cores in a 3D numerical simulation of a magnetized, turbulent, and self-gravitating molecular cloud. Cores are identified at the density threshold levels of $10^{4}$ and $2.5 \\times 10^{4}$ cm$^{-3}$. In agreement with observations, the slope of the core mass function is found to be steeper for cores identified at the higher threshold level. The steepening of the slope with increasing threshold is due to the increasing importance of gravity in the virial balance of the cores at higher densities and a sign of the role played by gravity in the gravo-turbulent scenario of star formation. On the other hand, at lower thresholds, the slope is shallower than what is found for the purely turbulent case. This is due to the effects of mass growth that affect the cores. This mass growth is essentially due in the initial phase to the effect of core coalescence which generates more massive structures particularly for cores defined at lower thresholds since they have larger cross sections." }, "0801/0801.0064_arXiv.txt": { "abstract": "{We present results from a kinematical study of the gas in the nucleus of a sample of three LINER galaxies, obtained from archival HST/STIS long-slit spectra. We found that, while for the elliptical galaxy NGC 5077, the observed velocity curves are consistent with gas in regular rotation around the galaxy's center, this is not the case for the two remaining objects. By modeling the surface brightness distribution and rotation curve from the emission lines in NGC 5077, we found that the observed kinematics of the circumnuclear gas can be accurately reproduced by adding to the stellar mass component a black hole mass of $M_{\\rm BH} = 6.8_{-2.8}^{+4.3}\\times 10^8 M_{\\odot}$ (uncertainties at a 1$\\sigma$ level); the radius of its sphere of influence ($R_{\\rm sph} \\sim$ 0\\farcs34) is well-resolved at the HST resolution. The BH mass estimate in NGC 5077 is in fairly good agreement with both the $M_{\\rm BH}-M_{\\rm bul}$ (with an upward scatter of $\\sim$ 0.4 dex) and $M_{\\rm BH}-\\sigma$ correlations (with an upward scatter of 0.5 dex in the Tremaine et al. form and essentially no scatter using the Ferrarese et al. form) and provides further support for the presence of a connection between the {\\sl residuals} \\rm from the $M_{\\rm BH}-\\sigma$ correlation and the bulge effective radius. This indicates the presence of a black hole's ``fundamental plane'' in the sense that a combination of at least $\\sigma$ and $R_{\\rm e}$ drives the correlations between $M_{\\rm BH}$ and host bulge properties. ", "introduction": "\\label{intro} It is now clear that the presence of a supermassive black hole (SMBH) is a common, if not universal, feature in the center of galaxies. In fact, since active galactic nuclei (AGNs) are thought to be powered by mass accretion onto an SMBH, the high incidence of low-luminosity AGN activity in nearby galaxies \\citep{heckman80, maoz95, ho97a, ho97b, braatz97, barth98, barth99,nagar02} has led to the conclusion that a significant fraction of galaxies in the Local Universe must host SMBH. This conclusion is now supported by direct measurements of SMBH masses in the centers of nearby galaxies obtained with different techniques \\citep[see][for a review]{ferrarese05}. These measurements indicate that the BH mass $M_{\\rm BH}$ is related to the properties of the host galaxy, such as bulge luminosity $L_{\\rm bul}$ and mass $M_{\\rm bul}$ \\citep{kormendy95, magorrian98, marconi03}, light concentration \\citep{graham01, graham07}, and bulge velocity dispersion $\\sigma_{\\rm bul}$ \\citep{ferrarese00, gebhardt00a, tremaine02}. The existence of any correlations between $M_{\\rm BH}$ and host bulge properties supports the idea that the growth of SMBHs and the formation of bulges are closely linked \\citep{silk98, haehnelt00}. This has profound implications for the process of galaxy formation and evolution. Moreover, SMBH mass estimates inferred via the above correlations, when more direct methods are not feasible, enter into a variety of important studies spanning from AGNs physics to the coeval formation and evolution of the host galaxy and its nuclear black hole. These results need, however, to be further investigated by increasing the number of accurate BH mass determinations in nearby galactic nuclei to set these correlations on a stronger statistical basis. In particular, such a study has the potential of establishing the precise role of the various host galaxy's parameters in setting the resulting BH mass. To date, reliable SMBH detections have been obtained for a limited number of galaxies ($\\sim$ 30, \\citealt{ferrarese05}), with the bulk of $M_{\\rm BH}$ estimates in the range of $10^{7}-10^{9} M_{\\odot}$. To add reliable new points to the $M_{\\rm BH}-$ host galaxy's properties planes is then a fundamental task for future developments of astronomical and physical studies. One widely applicable and relatively simple method of detecting BHs is based on gas kinematics (e.g. \\citealt{harms94,ferrarese96, macchetto97, barth01}), through studies of emission lines from circumnuclear gas disks, provided that the gas velocity field is not significantly influenced by non gravitational motions. However, the purely gravitational kinematics of the gas can be established {\\sl a posteriori} from successful modeling of the gas velocity field under the sole influence of the stellar and black hole potential. The Space Telescope Imaging Spectrograph (STIS) onboard $\\it {HST}$ is still the most suitable instrument for such studies as it provides a high angular resolution ($\\sim 0\\farcs1$) in the optical spectral region. In this band brighter emission lines are found with respect to the infrared, the only band accessible at high resolution with ground-based adaptive optics telescopes. The wealth of unpublished data contained in STIS archives represents an extraordinary and still unexplored resource. With the aim of finding galaxy candidates to provide a successful SMBH mass measurement, we performed a systematic search for unpublished data in the $\\it {HST}$ STIS archive. In this paper we present the results obtained for a sample of three LINER galaxies. The galaxies were observed with STIS on {$\\it HST$} during Cycle 7 under Proposal ID 7354 and were part of a larger sample of eight selected objects. The results on NGC 3998 have already been published \\citep[][hereafter Paper I]{ngc3998}. In a forthcoming paper, we will complete the proposal targets and discuss the relation between BH mass, host galaxy properties, and nuclear activity. Table \\ref{source} lists the galaxies and their main physical properties. Basic data are from the Lyon/Meudon Extragalactic Database (HyperLeda) or NASA/IPAC Extragalactic Database (NED). Distances are calculated with ${\\it H}_0$ = 75 km s$^{-1}$ Mpc$^{-1}$ and are corrected for Local Group infall onto Virgo. The paper is organized as follows. In Sec. \\ref{obs} we present HST/STIS data and the reduction that lead to the results described in Sec. \\ref{results}. In Sec. \\ref{fitting} we model the observed emission-line rotation curve for NGC 5077 and show that the dynamics of the circumnuclear gas can be accurately reproduced by circular motions in a thin disk when a point-like dark mass is added to the stellar potential. Our results are discussed in Sec. \\ref{discussion}, and summarized in Sec. \\ref{summary}. ", "conclusions": "\\label{summary} We have presented results from a gas kinematics study in the nucleus of three nearby LINER galaxies: IC 989, NGC 5077, and NGC 6500 using archival HST/STIS spectra. Only in the case of NGC 5077 does the nuclear velocity curves appear to be associated with gas in regular rotation. For IC 989 our results indicate an inner counter-rotating gas system, while for NGC 6500 the complex trend in the velocity curves suggests a nuclear expanding bubble. We used our modeling code to fit the observed [N II]$\\lambda$6583 surface brightness distribution and velocity curve of NGC 5077. The dynamics of the rotating gas can be accurately reproduced by motions in a thin disk when a compact dark mass of $M_{\\rm BH} = 6.8_{-2.8}^{+4.3}\\times 10^8 M_{\\odot}$ is added to the stellar mass component. Furthermore, the black hole in NGC 5077 has a sphere of influence radius, $R_{\\rm sph} = GM_{\\rm BH}/\\sigma_{\\rm star}^{2}$, of $\\sim$ 62 pc ($\\simeq$ $0\\farcs34$), well-resolved at the HST resolution (2$R_{\\rm sph}/R_{\\rm res} \\simeq$ 6.8). For what concerns the connections of this BH mass estimate with the properties of the host galaxy, the $M_{\\rm BH}$ value for NGC 5077 is in good agreement (within a factor of 2.3) with the $M_{\\rm BH}-M_{\\rm bul}$ correlation between BH and host bulge mass. The black hole mass predicted by the $M_{\\rm BH}-\\sigma$ correlation is a factor of 3.4 lower than our measure, adopting the relation found by \\citet{tremaine02} and in excellent agreement using the parameterization by \\citet{ferrarese05}. This result, in conjunction with the previous results for NGC 3998 (Paper I), strengthens the possibility of a connection between the residuals from the $M_{\\rm BH}-\\sigma$ relation and the bulge effective radius. While NGC 3998, indeed, has one of the lowest values of $R_{\\rm e}$ among galaxies with measured $M_{\\rm BH}$ and shows a negative residual, NGC 5077 has a larger effective radius and shows a small positive residual. We also recently showed that the same result was found for the Seyfert galaxy NGC 5252: a larger effective radius corresponds in this galaxy to a still larger positive residual. Apparently, only with a combination of at least $\\sigma$ and $R_{\\rm e}$ is it possible to account for the correlations between $M_{\\rm BH}$ and other bulge properties, indicating the presence of a black hole's ``fundamental plane''. Clearly, the number of direct black-hole mass measurements must be further increased, together with precise determinations of $\\sigma$ and $R_{\\rm e}$, to test these conclusions on a stronger statistical basis. In a forthcoming paper we will present new BH mass determinations and discuss the physical implications of our results." }, "0801/0801.0855_arXiv.txt": { "abstract": "In this Letter, we propose that a microphysical process takes a vital role in the shocked region in which the prompt emission of GRBs is emitted. The turbulent energy is included in the internal energy transferred by the kinetic energy of the shock. It dissipates through stochastic acceleration for the electrons to supply the early X-ray emission in the phase of shallow decay. We put the constraints on the time evolution of microphysical parameters. The early X-ray fluxes can be obtained by this scenario and these results are consistent with the Swift observation. ", "introduction": "In the Swift era \\citep{gehrels04}, the X-ray telescope (XRT) observation \\citep{burrows05} provides the complete light curves of gamma-ray bursts (GRBs) in the 0.2-10 keV band. One of the most interesting discoveries is the so-called shallow decay segment: the flux plateau of $F\\propto t^{-0.5}$ within $10^3-10^4$ second after the trigger \\citep{campana05,depasquale06}. Recent statistic analyses \\citep{nousek06,obrien06,liang07} have revealed that the phase of shallow decay in the X-ray afterglow might be a common feature of the long GRBs. The shallow decay in the early X-ray light curve is still a mystery, although theoretical explanations have been put forward from several aspects (see Zhang 2007 for a comprehensive review). Most of the models are: hydrodynamics of the shock by energy injection \\citep{granot06a,zhang06}, geometry of the jet \\citep{eichler06,toma06}, varying microphysical parameters \\citep{fan06,granot06b,panaitescu06,ioka06}, late prompt emission \\citep{ghisellini07} or up-scattered forward-shock emission \\citep{panaitescu07}. From the point of microphysics in the hydrodynamic evolution, the nature of coupling between the electrons, protons and magnetic field is complex \\citep{chiang99}. Usually the simple way is to assume the equipartition between electrons, protons and magnetic field \\citep{panaitescu98}. However, in this Letter, we propose that the kinetic energy of the relativistic shocks in the plasma has been converted to the internal energy as three parts: (1) $\\varepsilon_B$, which is the energy of magnetic field; (2) $\\varepsilon_e$ and $\\varepsilon_p$, it means that the electrons and protons/positrons are accelerated by the shocks, normally, this is the process of first-order Fermi acceleration; (3) $\\varepsilon_t$, which presents the turbulent energy, and this energy would sustain a relatively long time (see Section 2). The last part has not been taken into account by the former research. In our novel scenario, the relativistic electrons accelerated by the first-order Fermi acceleration emit the gamma-ray by synchrotron radiation, after the decrease of the prompt emission tail which is shown as the deep decay in the early X-ray light curve, indicating that the internal shocks are abated, the follow-up turbulence and its effects might be dominated. The turbulence could transfer its energy to the electrons via second-order Fermi acceleration, which is also called as the stochastic acceleration. Therefore, in this turbulent region, due to the resonant interaction between electrons and plasma waves, the electrons buried in the magnetic field are re-accelerated by the stochastic acceleration. The emission of these electrons dominates the shallow decay phase, until the turbulent energy dissipates and the external shock sweeps the surround medium thus the deep decay appears again. In Section 2, we review the Fokker-Planck equation and list the coefficients associated with the turbulent term. In Section 3, the turbulent parameter $\\varepsilon_t$, as same as $\\varepsilon_e$ and $\\varepsilon_B$, is introduced. Due to the turbulence, these microphysical parameters evolved with time are constrained by the process of stochastic acceleration. Finally, we select these relations to reproduce the feature of shallow decay. The discussions are given in Section 4. ", "conclusions": "Due to the turbulent energy and its dissipation in the shocked region, after the gamma-ray radiation, the electrons are re-accelerated by the stochastic acceleration and emit the early X-ray fluxes. Therefore, the energy injection in the central engine is not required to reproduce the emission in the shallow decay phase. Our results manifest that the temporal feature in the shallow decay phase represents the process of turbulent dissipation. The microphysical parameters, $\\varepsilon_e$, $\\varepsilon_B$ and $\\varepsilon_t$, are varied with the time in the shocked region. Given the special values of $p=2.0$ and $a=0.5$ in the adiabatic case, our calculation can represent the typical observational flux as $F\\propto t^{-0.5}$. Moreover, it is noted that the same shallow decay phase is also detected in the optical band as well \\citep{mason06}, our simple interpretation is that the synchrotron emission of X-ray band and optical band may be original from the same shocked region with turbulent energy. Another advantage of our model is that the crisis of radiative efficiency \\citep{lloyd04,ioka06,fan06,zhang07} is therefore dispelled without any additional assumption of ejection/ejecta from the central engine. From this point of view, we support the internal shock pattern of standard fireball model. In this work, we assume $f(\\gamma)\\propto \\gamma^{-p}$ where $p=2.2$ as the typical value. The detailed calculation of the energy distribution $f(\\gamma)$ is not needed, because in our scenario the timescale of acceleration is much smaller than that of turbulent energy dissipation. Since the timescale of acceleration is also smaller than that of shock hydrodynamics, therefore, during the phase of shallow decay, the index of spectrum does not change \\citep{liang07}. Generally, the varied values of $\\varepsilon_e$ and $\\varepsilon_B$ lead to the different hydrodynamic evolution and emissions in the entire afterglow \\citep{mao01a,mao01b}. In this Letter, the early X-ray fluxes in the shallow decay phase are estimated either in the adiabatic or radiative case through the turbulent process. The average slope of radiative case is deeper than that of adiabatic case. Compared with the XRT observation, the hydrodynamic evolution of adiabatic case could be the realized regime during the shallow decay phase. Although it is hard to obtain the values of $\\varepsilon_e$ and $\\varepsilon_B$ in our scenario, we put the constraints of $\\varepsilon_e\\ll 1$ and $\\varepsilon_B\\ll 1$. More observational samples are particularly requested for the further investigations." }, "0801/0801.3856_arXiv.txt": { "abstract": "We present photometry and spectroscopy of the suspected cataclysmic variable (CV) Lanning 386. We confirm that it is a CV, and observe deep eclipses, from which we determine the orbital period $P_{\\rm orb}$ to be $0.1640517 \\pm 0.0000001$ d (= 3.94 h). Photometric monitoring over two observing seasons shows a very active system with frequent outbursts of variable amplitude, up to $\\sim 2$ mag. The spectrum in quiescence is typical of dwarf novae, but in its high state the system shows strong \\ion{He}{2} emission and a broad \\ion{C}{4} Wolf-Rayet feature. This is unusual for dwarf novae in outburst and indicates a high excitation. In its high state the system shows some features reminiscent of an SW Sextantis-type CV, but lacks others. We discuss the classification of this puzzling object. ", "introduction": "Cataclysmic variables (CVs) are close binary stars in which a white dwarf primary accretes matter via Roche lobe overflow from a secondary, which usually resembles a lower main-sequence star \\citep{warner}. In most CVs an accretion disk encircles the white dwarf. Optical emission can come from several locations in the system. The gas stream from the secondary star strikes the outer rim of the accretion disk producing a rapidly fluctuating bright spot (or hot spot). The accretion disk often dominates the light, and instabilities in the disk can produce large variations in the system brightness. Strong white-dwarf magnetic fields can truncate or even entirely disrupt the accretion disk, allowing material to fall directly onto the white dwarf. In many systems the accumulated hydrogen-rich material can undergo a thermonuclear runaway, creating a classical nova outburst. The range of phenomena seen in CVs means that determining the subclass of any example can require extensive photometric and spectroscopic study. H. H. Lanning (\\citealt{lanning1973} and subsequent papers) conducted a search for ultraviolet-bright (UV) sources in the galactic plane by visually examining plates taken with the Palomar Oschin Schmidt. To date, six supplements have been published, revealing a total of 724 UV sources. Six of these are included in the {\\it Catalog and Atlas of Cataclysmic Variables} \\citep{downes2001} as confirmed or suspected CVs; they are listed in Table \\ref{lanningtable}. \\citet{eracleous2002} presented a spectrum of Lanning 386 which showed apparent CV features -- a blue continuum, He I absorption lines, and broad Balmer emission lines. The extensive photometric and spectroscopic observations we present here establish that Lanning~386 is indeed a deeply-eclipsing CV with $P_{\\rm orb} = 0.1640517 \\pm 0.0000001$ d. Intriguingly, it shares characteristics of both the dwarf novae -- CVs that undergo occasional eruptions thought to be caused by a disk instability -- and the SW Sextantis stars \\citep{thorstensen1991}, a subclass of the persistently-bright novalike variables. ", "conclusions": "Our photometry shows Lanning~386 to be an active, eclipsing CV, with $P_{\\rm orb} = 3.94$ h and an eclipse depth of $\\sim 1.5$ mag. Monitoring over many months shows bright outbursts reaching $V\\approx 15.2$ from a quiescent state with $V\\approx 17.2$. Lanning 386 does not fit neatly into any CV subclass. The long-term variability resembles that of a dwarf nova, and at minimum light the spectrum is typical of that class. However, when most dwarf novae outburst, their Balmer emission becomes very weak, and broad absorption wings appear (e.g., \\citealt{thorstensen98}); it is also unusual (though not unknown) for \\ion{He}{2} to become prominent (e.g., \\citealt{hessman84}). The high-state spectrum seen here is very much unlike this, resembling instead that of a novalike variable. The \\ion{He}{1} line profiles, and the unusual strength of \\ion{He}{2}, are reminiscent of an SW Sextantis star \\citep{thorstensen1991}. In addition, Lanning 386 is similar to other SW Sex stars in that the emission lines are eclipsed much more shallowly than the continuum, indicating a spatially extended source. The rotational disturbance around eclipse is also seen in some other SW Sex stars (e.g., \\citealt{rodriguez04}). However, other canonical SW Sex characteristics are missing here. The emission line radial-velocity phase in SW Sex stars lags far behind (by $> 45^{\\circ}$) the phase expected for the white dwarf motion, based on the eclipse ephemeris, but in Lanning 386 the phase lag is only $\\sim 11^\\circ \\pm 2^\\circ$. In addition, while the \\ion{He}{1} lines do appear in absorption for part of the orbit, in other SW Sex stars the absorption appears near phase 0.5, whereas it appears somewhat later here. Nonetheless, the unusual maximum-light spectrum suggests that we should consider carefully whether the long-term light curve in Fig.~\\ref{detailplot} forces us to classify this star as a dwarf nova. Might another mechanism explain the variability in a manner more consistent with the maximum-light spectrum? Many novalike variables, and many SW Sex stars in particular, are also VY Scl stars -- novalike variables that occasionally fade deeply for varying amounts of time, for reasons that are not entirely understood. Might Lanning 386 be a VY Scl star masquerading as a dwarf nova? This seems unlikely, for two reasons. First, the long-term light curves of VY Scl stars tend to look quite different from those of dwarf novae -- VY Scl stars typically remain in a high state for months or years, and only occasionally fade by large amounts, up to 5 mag (see examples in \\citealt{honeycutt04}). The light curve here, by contrast, consists of frequent, moderate-amplitude outbursts from a faint state. Second, the spectra of VY Scl stars in their faint states look as if accretion has nearly stopped, with the Balmer emission lines typically becoming very narrow and the underlying stars becoming evident \\citep{shafter1983, schneider81}. This is very different from the essentially normal dwarf nova spectrum seen here. The DQ Her stars, also called intermediate polars, are magnetic systems \\citep[for a review, see][]{patterson94}; some of these undergo outbursts that at least superficially resemble dwarf nova outbursts. Typically, DQ Her outbursts are infrequent compared to those of most dwarf novae; they also tend to decay rapidly \\citep{hellier97}. For this reason, there is little spectroscopy available for outbursts of these systems. However, spectra do exist for EX Hya in outburst, which, show strong emission lines with interesting velocity structure, reminiscent of the SW Sex stars \\citep{hellier89,hellier00}. In this way, EX Hya resembles Lanning 386, but the morphological fit is imperfect -- the brightenings we see in Lanning 386 appear to be more frequent than those seen in most DQ Her stars. Also, we have (as of yet) no direct evidence that Lanning 386 is a magnetic system, because we are unaware of any searches for the coherent pulsations and/or circular polarization that would prove the case. It has been suggested that the outbursts seen in some magnetic systems are due to mass-transfer bursts rather than the disk instabilities that are thought to cause normal dwarf nova outbursts \\footnote{It is possible that mass-transfer bursts could be triggered by an event in the disk.}. During a period of increased mass transfer, material overflowing the disk could give rise to enhanced emission -- there is evidence for this in EX Hya \\citep{hellier00}. It may be that the outbursts seen in Lanning 386, whatever their trigger, involve enhanced mass transfer -- which may or may not be exclusive to magnetic systems. We can summarize the morphological conundrum presented by Lanning 386 as follows: Its minimum-light spectrum and the light curve suggest it is a dwarf nova with frequent, relatively low-amplitude outbursts, while at maximum light it resembles a novalike variable, with some (but not all) the characteristics of an SW Sex star. EX Hya, a magnetic system, shows a similar set of symptoms, but the presentation is not identical." }, "0801/0801.1548_arXiv.txt": { "abstract": "The residuals about the standard $M_{\\rm bh}$-$\\sigma$ relation correlate with the effective radius, absolute magnitude, and S\\'ersic index of the host bulge. Although, it is noted here that the elliptical galaxies do not partake in such correlations. Moreover, it is revealed that barred galaxies (with their relatively small, faint, and low stellar concentration bulges) can deviate from the $M_{\\rm bh}$-$\\sigma$ relation by $\\delta \\log M_{\\rm bh} \\approx -0.5$ to $-1.0$ dex (their $\\sigma$ values are too large) and generate much of the aforementioned correlations. Removal of the seven barred galaxies from the Tremaine et al.\\ set of 31 galaxies gives a ``barless $M_{\\rm bh}$-$\\sigma$'' relation with an intrinsic scatter of 0.17 dex (cf.\\ 0.27 dex for the 31 galaxies) and a total scatter of 0.25 dex (cf.\\ 0.34 dex for the 31 galaxies). The introduction of a third parameter does not reduce the scatter. Furthermore, removal of the barred galaxies, or all the disk galaxies, from an expanded and updated set of 40 galaxies with direct black hole mass measurements gives a consistent result, such that $\\log(M_{\\rm bh}/M_{\\sun}) = (8.25\\pm0.05) + (3.68\\pm0.25)\\log [\\sigma/200\\, {\\rm km\\, s}^{-1}]$. The (barless) $\\sigma$-$L$ relation for galaxies with black hole mass measurements is found to be consistent with that from the SDSS sample of early-type galaxies. In addition the barless $M_{\\rm bh}$-$\\sigma$ relation, the $M_{\\rm bh}$-$n$ relation, and the $M_{\\rm bh}$-$L$ relation are shown to yield SMBH masses less than 2-4$\\times 10^9 M_{\\sun}$. ", "introduction": "Tight correlations between supermassive black hole (SMBH) masses and large scale properties of the host bulges are interesting for two obvious reasons. They enable us to predict SMBH masses in thousands of galaxies where the black hole's sphere-of-influence is highly unresolved, and they provide clues to the physical processes responsible for the co-evolution of black hole and host bulge. Recent endeavors have advocated relations involving not one but two bulge parameters and therein claims of ``fundamental planes'', akin to the Fundamental Plane for elliptical galaxies (Djorgovski \\& Davis 1987; Dressler et al.\\ 1987). The existence of such SMBH fundamental planes imply that current theories for the $M_{\\rm bh}$-$\\sigma$ relation (Ferrarese \\& Merritt 2000; Gebhardt et al.\\ 2000) or the $M_{\\rm bh}$-$L$ relation (McLure \\& Dunlop 2002; Marconi \\& Hunt 2003; updated in Graham 2007), which do not include a third parameter, are incomplete. This article investigates the fundamental planes for SMBHs\\footnote{The ``fundamental plane of black hole activity'' involving radio core luminosity and X-ray luminosity (Merloni et al.\\ 2003; Falcke et al.\\ 2004) is not addressed here.} involving the parameters $M_{\\rm bh}$, $R_{\\rm e}$, and $\\langle \\mu \\rangle_{\\rm e}$ (Barway \\& Kembhavi 2007), $M_{\\rm bh}$, $R_{\\rm e}$, and $\\sigma$ (Marconi \\& Hunt 2003; de Francesco et al.\\ 2006; Aller \\& Richstone 2007; Hopkins et al.\\ 2007), and, for the first time, $M_{\\rm bh}$, $\\sigma$, and $n$. In Section~2 it is explained why a previous claim for a small `total' scatter (0.19 dex) about the $M_{\\rm bh}$-$R_{\\rm e}$-$\\langle \\mu \\rangle _{\\rm e}$ plane was the result of a miscalculation. In Section~3 it is revealed that the galaxies which deviate from the $M_{\\rm bh}$-$\\sigma$ relation, giving rise to the $M_{\\rm bh}$-$\\sigma$-$R_{\\rm e}$, $M_{\\rm bh}$-$\\sigma$-$L$, and $M_{\\rm bh}$-$\\sigma$-$n$ relations with less scatter than the $M_{\\rm bh}$-$\\sigma$ relation, are predominantly barred galaxies. A ``barless $M_{\\rm bh}$-$\\sigma$'' relation, and an elliptical-only $M_{\\rm bh}$-$\\sigma$ relation, is subsequently constructed in Section~4 and found to heavily nullify the evidence for fundamental planes for SMBHs and their host bulges. Given the recent discussion in the literature about biases in the $M_{\\rm bh}$-$\\sigma$ and/or $M_{\\rm bh}$-$L$ relation, and also in the local sample of galaxies with direct SMBH mass measurements, these concerns are explored here. In Sections~5 a $\\sigma$-$L$ relation is constructed and shown to be equal to that obtained using SDSS early-type galaxy data, thereby laying to rest concerns that the local sample of galaxies with direct SMBH mass measurements may be biased with respect to the greater population (e.g.\\ Yu \\& Tremaine 2002; Bernardi et al.\\ 2007). Furthermore, in section~6, the $K$-band $M_{\\rm bh}$-$L$ relation and the barless $M_{\\rm bh}$-$\\sigma$ relation are shown to yield consistent results with neither giving SMBH masses greater than $\\sim4 \\times 10^9 M_{\\sun}$. ", "conclusions": "" }, "0801/0801.1062_arXiv.txt": { "abstract": "The study of the metal enrichment of the intra--cluster and inter--galactic media (ICM and IGM) represents a direct means to reconstruct the past history of star formation, the role of feedback processes and the gas--dynamical processes which determine the evolution of the cosmic baryons. In this paper we review the approaches that have been followed so far to model the enrichment of the ICM in a cosmological context. While our presentation will be focused on the role played by hydrodynamical simulations, we will also discuss other approaches based on semi--analytical models of galaxy formation, also critically discussing pros and cons of the different methods. We will first review the concept of the model of chemical evolution to be implemented in any chemo--dynamical description. We will emphasise how the predictions of this model critically depend on the choice of the stellar initial mass function, on the stellar life--times and on the stellar yields. We will then overview the comparisons presented so far between X--ray observations of the ICM enrichment and model predictions. We will show how the most recent chemo--dynamical models are able to capture the basic features of the observed metal content of the ICM and its evolution. We will conclude by highlighting the open questions in this study and the direction of improvements for cosmological chemo--dynamical models of the next generation. ", "introduction": "\\label{Introduction} Clusters of galaxies are the ideal cosmological signposts to trace the past history of the inter--galactic medium (IGM), thanks to the high density and temperature reached by the cosmic baryons trapped in their gravitational potential wells (\\citealt{2002ARA&A..40..539R}; \\citealt{2005RvMP...77..207V}; \\citealt{diaferio2008} - Chapter 2, this volume). Observations in the X--ray band with the Chandra and XMM--Newton satellites are providing invaluable information on the thermodynamical properties of the intra--cluster medium (ICM) (\\citealt{kaastra2008} - Chapter 9, this volume). These observations highlight that non--gravitational sources of energy, such as energy feedback from supernovae (SNe) and Active Galactic Nuclei (AGN) have played an important role in determining the ICM physical properties. Spatially resolved X--ray spectroscopy permits to measure the equivalent width of emission lines associated to transitions of heavily ionised elements and, therefore, to trace the pattern of chemical enrichment (e.g., \\citealt{2004cgpc.symp..123M} for a review). In turn, this information is inextricably linked to the history of formation and evolution of the galaxy population (e.g., \\citealt{1997ApJ...488...35R} and references therein), as inferred from observations in the optical/near-IR band. For instance, \\citet{2004A&A...419....7D} have first shown that cool core clusters are characterised by a significant central enhancement of the iron abundance, which closely correlates with the magnitude of the Brightest Cluster Galaxies (BCGs). This demonstrates that a fully coherent description of the evolution of cosmic baryons in the condensed stellar phase and in the diffuse hot phase requires properly accounting for the mechanisms of production and release of both energy and metals. The study of how these processes take place during the hierarchical build up of cosmic structures has been tackled with different approaches. Semi--analytical models (SAMs) of galaxy formation provide a flexible tool to explore the space of parameters which describe a number of dynamical and astrophysical processes. In their most recent formulation, such models are coupled to dark matter (DM) cosmological simulations, to trace the merging history of the halos where galaxy formation takes place, and include a treatment of metal production from type-Ia and type-II supernovae (SN\\,Ia and SN\\,II, hereafter; \\citealt{2004MNRAS.349.1101D}; \\citealt{2005MNRAS.358.1247N}), so as to properly address the study of the chemical enrichment of the ICM. The main limitation of this method is that it does not include the gas dynamical processes which causes metals, once produced, to be transported in the ICM. As a consequence, they provide a description of the global metallicity of clusters and its evolution, but not of the details of its spatial distribution. In order to overcome this limitation, \\citet{2006MNRAS.368.1540C} applied an alternative approach, in which non--radiative hydrodynamical simulations of galaxy clusters are used to trace at the same time the formation history of DM halos and the dynamics of the gas. In this approach, metals are produced by SAM galaxies and then suitably assigned to gas particles, thereby providing a chemo--dynamical description of the ICM. \\citet{2006A&A...452..795D} used hydrodynamical simulations, which include prescriptions for gas cooling, star formation and feedback, similar to those applied in SAMs, to address the specific role played by ram--pressure stripping to distribute metals, while \\citet{2007A&A...466..813K} used the same approach to study the different roles played by galactic winds and by ram--pressure stripping. While these approaches offer advantages with respect to standard SAMs, they still do not provide a fully self--consistent picture, in which chemical enrichment is the outcome of the process of star formation, associated to the cooling of the gas infalling in high density regions. In this sense, a fully self--consistent approach requires that the simulations include the processes of gas cooling, star formation and evolution, along with the corresponding feedback in energy and metals. A number of authors have presented hydrodynamical simulations for the formation of cosmic structures, which include treatments of the chemical evolution at different levels of complexity. \\citet{1996A&A...315..105R} presented SPH simulations of the Galaxy, forming in an isolated halo, by following iron and oxygen production from SN\\,II and SN\\,Ia, also accounting for the effect of stellar lifetimes. \\citet{2001MNRAS.325...34M} performed a detailed analysis of chemo--dynamical SPH simulations, aimed at studying both numerical stability of the results and the enrichment properties of galactic objects in a cosmological context. \\citet{2002MNRAS.330..821L} discussed a statistical approach to follow metal production in SPH simulations, which have a large number of star particles, showing applications to simulations of a disc--like galaxy and of a galaxy cluster. \\citet{2003MNRAS.346..135K} carried out cosmological chemo--dynamical simulations of elliptical galaxies, with an SPH code, by including the contribution from SN\\,Ia and SN\\,II, also accounting for stellar lifetimes. \\citet{2003MNRAS.339.1117V} performed an extended set of cluster simulations and showed that profiles of the iron abundance are steeper than the observed ones. A similar analysis has been presented by \\citet{2006MNRAS.371..548R}, who also considered the effect of varying the IMF and the feedback efficiency on the enrichment pattern of the ICM. \\citet{2005MNRAS.364..552S} presented an implementation of a model of chemical enrichment in the {\\tt GADGET-2} code, coupled to a self--consistent model for star formation and feedback. In their model, which was applied to study the enrichment of galaxies, they included the contribution from SN\\,Ia and SN\\,II, assuming that all long--lived stars die after a fixed delay time. \\citet{Tornatore07,2004MNRAS.349L..19T} presented results from an implementation of a detailed model of chemical evolution model in the {\\tt GADGET-2} code \\citep{2005MNRAS.364.1105S}, including metallicity--dependent yields and the contribution from intermediate and low mass stars (ILMS hereafter). The major advantage of this approach is that the metal production is self--consistently predicted from the rate of gas cooling treated by the hydrodynamical simulations, without resorting to any external model. However, at present the physical scales involved by the processes of star formation and SN explosions are far from being resolved in simulations which sample cosmological scales. For this reason, such simulations also need to resort to external sub--resolution models, which provide an effective description of a number of relevant astrophysical processes. The aim of this paper is to provide a review of the results obtained so far in the study of the chemical enrichment of the ICM in a cosmological context. Although we will concentrate the discussion on the results obtained from full hydrodynamical simulations, we will also present results based on SAMs. As such, this paper complements the reviews by \\citealt{borgani2008} - Chapter 13, this volume, which reviews the current status in the study of the thermodynamical properties of the ICM with cosmological hydrodynamical simulations, and by \\citealt{schindler2008} - Chapter 17, this volume, which will focus on the study of the role played by different physical processes in determining the ICM enrichment pattern. Also, this paper will not present a detailed description of the techniques for simulations of galaxy clusters, which is reviewed by \\citealt{10_dolag2008} - Chapter 12, this volume. In Sect.~\\ref{chemical_evolution} we review the concept of model of chemical evolution and highlight the main quantities which are required to fully specify this model. In Sect.~\\ref{globab} we review the results on the global metal content of the ICM, while Sect.~\\ref{profiles} concentrates on the study of the metallicity profiles and Sect.~\\ref{evol} on the study of the ICM evolution. Sect.~\\ref{galaxies} discusses the properties of the galaxy population. Finally, Sect.~\\ref{summary} provides a critical summary of the results presented, by highlighting the open problems and lines of developments to be followed by simulations of the next generation. ", "conclusions": "\\label{summary} The study of the enrichment pattern of the ICM and its evolution provides a unique means to trace the past history of star formation and the gas--dynamical processes, which determine the evolution of the cosmic baryons. In this overview, we discussed the concept of model of chemical evolution, which represents the pillar of any chemo--dynamical model, and presented the results obtained so far in the literature, based on different approaches. Restricting the discussion to those methods which follow the chemical enrichment during the cosmological assembly of galaxies and clusters of galaxies, they can be summarised in the following categories.\\\\ {\\bf (1)} Semi--analytical models (SAMs) of galaxy formation \\citep{2004MNRAS.349.1101D,2005MNRAS.358.1247N}: the production of metals is traced by the history of galaxy formation within the evolving DM halos and no explicit gas--dynamical description of the ICM is included. These approaches are computationally very cheap and provide a flexible tool to explore the space of parameters determining galaxy formation and chemical evolution. A limitation of these approaches is that, while they provide predictions on the global metal content of the ICM, the absence of any explicit gas dynamical description causes the lack of useful information on the spatial distribution of metals.\\\\ {\\bf (2)} SAMs coupled to hydrodynamical simulations. In this case, the galaxy formation is followed as in the previous approach, but the coupling with a hydrodynamical simulation allows one to trace the fate of the enriched gas and, therefore, to study the spatial distribution of metals. Different authors have applied different implementations of this hybrid approach, by focusing either on the role of the chemical evolution model \\citep{2006MNRAS.368.1540C} or on the effect of specific gas--dynamical processes, such as ram--pressure stripping and galactic winds, in enriching the ICM (e.g., \\citealt{2007A&A...466..813K}). The advantage of this approach is clearly that gas--dynamical processes are now included at some level. However, the galaxy formation process is not followed in a self--consistent way from the cooling of the gas during the cosmic evolution.\\\\ {\\bf (3)} Full hydrodynamical simulations, which self--consistently follow gas cooling, star formation and chemical evolution \\citep{2003MNRAS.339.1117V,2003MNRAS.346..135K,2005MNRAS.361..983R,Tornatore07}. Rather than being described by an external recipe, galaxy formation is now the result of the cooling and of the conversion of cold dense gas into stars, as implemented in the simulation code. The major advantage of this approach is that the enrichment process now comes as the result of the star formation predicted by the hydrodynamical simulation, without resorting to any external model. The bottleneck, in this case, is represented by the computational cost, which becomes prohibitively high if one wants to resolve the whole dynamic range covered by the processes of metal production and distribution, feedback energy release, etc. For this reason, these simulation codes also need to resort to sub--resolution models, which provide an effective description of a number of physical processes. However, there is no doubt that the ever improving supercomputing technology and code efficiency make, in perspective, this approach the way to study the past history of cosmic baryons. As already emphasised, developing a reliable model of the cosmic history of metal production is a non--trivial task. This is due to the sensitive dependence of the model predictions on both the parameters defining the chemical evolution model (i.e., IMF, stellar life--times, stellar yields, etc.), and on the processes, such as ejecta from SN explosions and AGN, stripping, etc., which at the same time regulate star formation and transport metals outside galaxies. In the light of these difficulties, it is quite reassuring that the results of the most recent hydrodynamical simulations are in reasonable agreement with the most recent observational data from Chandra and XMM--Newton. In spite of this success, cluster simulations of the present generation still suffer from important shortcomings. The most important of them is probably represented by the difficulty in regulating overcooling at the cluster centre, so as to reproduce at the same time the observed ``cool cores'' and the passively evolving massive ellipticals. Besides improving the description of the relevant astrophysical processes in simulation codes, another important issue concerns understanding the possible observational biases which complicate any direct comparison between data and model predictions \\citep{2007arXiv0707.1573K,2007arXiv0707.2614R}. In this respect, simulations provide a potentially ideal tool to understand these biases. Mock X--ray observations of simulated clusters, which include the effect of instrumental response, can be analysed exactly in the same way as real observational data. The resulting ``observed'' metallicity can then be compared with the true one to calibrate out possible systematics. There is no doubt that observations and simulations of the chemical enrichment of the ICM should go hand in hand in order to fully exploit the wealth of information provided by X--ray telescopes of the present and of the next generation." }, "0801/0801.3251_arXiv.txt": { "abstract": "Tachyonic inflationary universe model in the context of a Chaplygin gas equation of state is studied. General conditions for this model to be realizable are discussed. By using an effective exponential potential we describe in great details the characteristic of the inflationary universe model. The parameters of the model are restricted by using recent astronomical observations. ", "introduction": "It is well known that inflation is to date the most compelling solution to many long-standing problems of the Big Bang model (horizon, flatness, monopoles, etc.) \\cite{guth,infla}. One of the success of the inflationary universe model is that it provides a causal interpretation of the origin of the observed anisotropy of the cosmic microwave background (CMB) radiation, and also the distribution of large scale structures \\cite{astro}. In concern to higher dimensional theories, implications of string/M-theory to Friedmann-Robertson-Walker (FRW) cosmological models have recently attracted great deal of attention, in particular, those related to brane-antibrane configurations such as space-like branes\\cite{sen1}. In recent times a great amount of work has been invested in studying the inflationary model with a tachyon field. The tachyon field associated with unstable D-branes might be responsible for cosmological inflation in the early evolution of the universe, due to tachyon condensation near the top of the effective scalar potential \\cite{sen2}, which could also add some new form of cosmological dark matter at late times \\cite{Sami_taq}. In fact, historically, as was empathized by Gibbons \\cite{gibbons}, if the tachyon condensate starts to roll down the potential with small initial velocity, then a universe dominated by this new form of matter will smoothly evolve from a phase of accelerated expansion (inflation) to an era dominated by a non-relativistic fluid, which could contribute to the dark matter detected in these days. On the other hand, the generalized Chaplygin gas has been proposed as an alternative model for describing the accelerating of the universe. The generalized Chaplygin gas is described by an exotic equation of state of the form \\cite{Bento} \\begin{equation} p_{ch} = - \\frac{A}{\\rho_{ch}^\\beta},\\label{1} \\end{equation} where $\\rho_{ch}$ and $p_{ch}$ are the energy density and pressure of the generalized Chaplygin gas, respectively. $\\beta$ is a constant that lies in the range $0 <\\beta\\leq 1$, and $A$ is a positive constant. The original Chaplygin gas corresponds to the case $\\beta = 1$ \\cite{2}. Inserting this equation of state into the relativistic energy conservation equation leads to an energy density given by \\cite{Bento} \\begin{equation} \\rho_{ch}=\\left[A+\\frac{B}{a^{3(1+\\beta)}}\\right]^{\\frac{1}{1+\\beta}}, \\label{2} \\end{equation} where $a$ is the scale factor and $B$ is a positive integration constant. The Chaplygin gas emerges as a effective fluid of a generalized d-brane in a (d+1, 1) space time, where the action can be written as a generalized Born-Infeld action \\cite{Bento}. These models have been extensively studied in the literature \\cite{other}. The model parameters were constrained using currents cosmological observations, such as, CMB \\cite{CMB} and supernova of type Ia (SNIa) \\cite{SIa}. In the model of Chaplygin inspired inflation usually the scalar field, which drives inflation, is the standard inflaton field, where the energy density given by Eq.(\\ref{2}), can be extrapolate for obtaining a successful inflation period with a Chaplygin gas model\\cite{Ic}. Recently, the dynamics of the early universe and the initial conditions for inflation in a model with radiation and a Chaplygin gas was studied in Ref.\\cite{Monerat:2007ud}. As far as we know, a Chaplygin inspired inflationary model in which a tachyonic field is considered has not been studied. The main goal of the present work is to investigate the possible realization of a Chaplygin inflationary universe model, where the energy density is driven by a tachyonic field. We use astronomical data for constraining the parameters appearing in this model. The outline of the paper is a follows. The next section presents a short review of the modified Friedmann equation by using a Chaplygin gas, and we present the tachyon-Chaplygin inflationary model. Section \\ref{sectpert} deals with the calculations of cosmological perturbations in general term. In Section \\ref{exemple} we use an exponential potential for obtaining explicit expression for the model. Finally, Sect.\\ref{conclu} summarizes our findings. ", "conclusions": "} In this paper we have studied the tachyon-Chaplygin inflationary model. In the slow-roll approximation we have found a general relation between the tachyonic potential and its derivative. This has led us to a general criterium for inflation to occur (see Eq.(\\ref{cond})). We have also obtained explicit expressions for the corresponding scalar spectrum index $n_s$ and its running $\\alpha_s$. By using an exponential potential with $\\alpha$ fixed (see Eq.(\\ref{pot})) and from the WMAP three year data, we found the values of the parameter $A$ and an upper limit for the tachyon potential $V_*$. In order to bring some explicit results we have taken $\\alpha=4\\times 10^{-6} m_p$ and $n_s=0.97$, from which we get the values $A\\simeq 2\\times10^{-23}m_p^8$, $V_*\\simeq 5\\times 10^{-13}m_p^4$ and $\\alpha_s\\simeq -0.02$. The restrictions imposed by currents observational data allowed us to establish a small range for the parameter $\\alpha$, which become $10^{-6}<\\alpha/m_p\\lesssim 10^{-5}$. From this range, and from Eqs.(\\ref{A}) and (\\ref{constrain}), we obtained the ranges $00.1h\\,$Mpc$^{-1}$) where linear biasing is not valid. This must be improved in the future: especially for low-redshift mass tracers a large fraction of the information is in the nonlinear regime, and if one aims for constraints at the several percent level, simple ideas like linear biasing may not be adequate even at $k=0.1h\\,$Mpc$^{-1}$. There is also uncertainty in the redshift distribution of the galaxies. One obvious method is to do lensing of the CMB in autocorrelation (with the four-point function \\cite{2001PhRvD..64h3005H} or using the iterative approaches \\cite{2003PhRvD..68h3002H}), although this will be very demanding on control of instrumental systematics. If one chooses to go the cross-correlation route, then when sufficiently high signal-to-noise ratio is available it will be desirable to use halo modeling, and to better calibrate the photometric redshifts for the large scale structure tracers with spectroscopic surveys. The latter will of course also be valuable to any cosmic shear program using the galaxies as lensing sources. \\end{list} If these challenges can be met, weak lensing of the CMB will make the transition from being a simple consistency check of the cosmological model to a routinely used cosmological probe." }, "0801/0801.2194_arXiv.txt": { "abstract": "We report a comprehensive statistical analysis of the observational data of the cosmic evolution of supernova (SN) rate density, to derive constraints on cosmic star formation history and the nature of type Ia supernova (SN Ia) progenitor. We use all available information of magnitude, SN type, and redshift information of both type Ia and core-collapse (CC) SNe in GOODS and SDF, as well as SN Ia rate densities reported in the literature. Furthermore, we also add 157 SN candidates in the past Subaru/Suprime-Cam data that are newly reported here, to increase the statistics. We find that the current data set of SN rate density evolution already gives a meaningful constraint on the evolution of the cosmic star formation rate (SFR) at $z \\lesssim 1$, though strong constraints cannot be derived for the delay time distribution (DTD) of SNe Ia. We derive a constraint of $\\alpha \\sim $ 3--4 [the evolutionary index of SFR density $\\propto (1+z)^\\alpha$ at $z \\lesssim 1$] with an evidence for a significant evolution of mean extinction of CC SNe [$E(B-V) \\sim 0.5$ at $z \\sim 0.5$ compared with $\\sim 0.2$ at $z=0$], which does not change significantly within a reasonable range of various DTD models. This result is nicely consistent with the systematic trend of $\\alpha$ estimates based on galactic SFR indicators in different wavelengths (ultraviolet, H$\\alpha$, and infrared), indicating that there is a strong evolution in mean extinction of star forming regions in galaxies at relatively low redshift range of $z \\lesssim 0.5$. These results are obtained by a method that is completely independent of galaxy surveys, and especially, there is no detection limit about the host galaxy luminosity in our analysis, giving a strong constraint on the star formation activity in high-$z$ dwarf galaxies or intergalactic space. ", "introduction": "In recent years a number of searches for high redshift supernovae (SNe) have been conducted. Although the primary purpose of most of these surveys is measurement of the cosmic expansion, these surveys also allowed measurements of the cosmic supernova rate density and its evolution \\citep{Pain02, Tonry03, Madgwick03, Gal-Yam04, Blanc04, Maoz04, Dahlen04, Cappellaro05, Barris06, Neill06, Poznanski07, Sharon07, Mannucci07, Kuznetsova07}. Studying these data should provide us with important information not only for the cosmic star formation history (CSFH) but also the still unknown progenitor of type Ia supernovae (SNe Ia). The progenitor of SNe Ia is believed to be a binary system including a white dwarf, and the SN Ia rate density evolution is a convolution of the cosmic star formation history and the delay time distribution (DTD) from star formation to SN Ia events. DTD depends on the progenitor models, and hence to constrain DTD observationally is a useful approach to reveal the SN Ia progenitor (Madau et al. 1998; Yungelson \\& Livio 1998, 2000; Dahlen \\& Fransson 1999; Gilliland et al. 1999; Gal-Yam \\& Maoz 2004; Strolger et al. 2004, 2005; Barris et al. 2004; Oda \\& Totani 2005, hereafter OT05; Strigari et al. 2005). However, it is not an easy task to actually extract useful constraints from the SN rate density evolution data. Previous studies \\citep{Maoz04, Strolger04, Forster06} mainly concentrated on the determination of DTD, using the rate density evolution of SNe Ia. In such an analysis, sometimes CSFH models are assumed based on the observational estimates from high-$z$ galaxy surveys. However, as argued by \\citet{Forster06}, the constraint on DTD models sensitively depends on the assumed CSFH, and hence it is difficult to derive a robust constraint on DTD. The primary purpose of this paper is to perform a comprehensive likelihood analysis using all available SN rate density evolution data in the literature, to derive constraints on DTD and/or CSFH. After the GOODS high-$z$ supernova survey \\citep{Dahlen04, Strolger04}, whose data was used in \\citet{Strolger04} and \\citet{Forster06}, a number of observational estimates of SN rate density evolution have been published (mostly for SNe Ia, but some data also for CC SNe). Our approach is to derive constraints only by using SN rate data, without using information of CSFH from galaxy surveys. We will perform a simultaneous fit to both the SN Ia and CC SN rate density evolution data, surveying parameters of the CSFH model with a variety of DTD models. We will find that, though a strong constraint on DTD models cannot be derived even from all the available data so far, we can set interesting constraints on CSFH and evolution of mean dust extinction of CC SNe, which can be compared with those inferred from galaxy surveys. Although there are a number of observational estimates on CSFH at a variety of redshifts from galaxy surveys, there is still a large uncertainty in the star formation rate (SFR) density estimated from galaxy observations, because of extinction, initial mass function, or extrapolation of luminosity functions to fainter magnitudes below the detection limits (see Hopkins 2004 and Hopkins \\& Beacom 2006, and reference therein). Therefore it is useful and important to derive constraints on CSFH from SNe independently of galaxy surveys. In contrast to SFR density estimates by galaxies, detectability of SNe does not depend on the host galaxy brightness, and even intergalactic star formation activity can be probed by hostless SNe. Searches for $z \\sim 1$ SNe are typically performed at wavelength around the $i'$ and $z'$ band roughly corresponding to the rest-frame visual bands, and hence the effect of extinction by dust is expected to be smaller than the CSFH estimates based on the rest-frame UV emission of galaxies. It is not trivial that a unit mass of star formation always produces the same number of SNe, but it could evolve with redshift or physical properties of galaxies. If a significant difference between CSFH inferred from galaxy surveys and that from SN surveys is found, it might indicate that the relation between star formation and supernova production is not as simple as normally assumed. In addition to the available SN rate density data in the literature, we also utilize the photometric sample of SN candidates found in the past observations using Subaru/Sprime-Cam. This Subaru Supernova Survey (SSS) sample includes 157 supernova candidates, 61 out of which have clear offsets from the centers of host galaxies and hence they are most likely SNe. This data set is complementary to GOODS, SNLS and the IfA deep survey \\citep{Strolger04,Neill06,Barris06} in terms of the combination of the survey area and depth; the covered area of SSS, $1.4$ deg$^2$, is wider than the GOODS, and the SSS depth, $i' \\sim 26.0$, is deeper than the SNLS and the IfA deep survey. Though no SN type or redshift information is available for the SSS sample, we add this data set to our likelihood analysis to increase the statistics especially for CC SNe. Compared with SNe Ia, there are not many data of the rate density for CC SNe. Combined analysis of the SSS counts including all SNe and other data for SN Ia rate density evolution should give some constraints on the CC SN rate density evolution and hence CSFH. The following are the plan of this paper. In \\S \\ref{sec:data} and \\S \\ref{sec:obs_res} we describe the SSS data set and analysis procedure of selecting SN candidates. Formulations of the comparison of the theoretical model and the observational data are given in \\S \\ref{sec:comp_method}. Constraints on the CSFH from our comprehensive parameter survey are derived in \\S \\ref{sec:results}. Conclusions are given in \\S \\ref{sec:summary}. Throughout this paper, the standard $\\Lambda$CDM universe is assumed with the following values of the cosmological parameters: $\\Omega_{\\rm M}$ = 0.3, $\\Omega_{\\rm \\Lambda}$ = 0.7, $h_{70} \\equiv H_{0}$/ (70 km s$^{-1}$ Mpc$^{-1}$) = 1. All magnitudes are given in the AB magnitude system. ", "conclusions": "\\label{sec:summary} We performed a comprehensive likelihood analysis of almost all available data of the cosmic SN rate density evolution both for CC and type Ia supernovae, to get information on the CSFH and the DTD of SNe Ia. We utilized the variability magnitude and redshift distribution of CC and Ia SNe of the GOODS and SDF-SNS, and other estimates of SN Ia rate density at various redshift in the literature. Furthermore, we added photometrically found supernova candidates in the past imaging data of Subaru/Suprime-Cam (Subaru Supernova Survey, SSS) to increase the statistics. The analysis of the SSS data is newly reported here, including 157 SN candidates down to $i' \\sim 26.0$ in the total survey area of 1.4 deg$^2$. 61 of the 157 SSS candidates are associated with host galaxies with significant offsets from galaxy centers, and hence they are almost certainly supernovae. Though the type and redshift information is not available for SSS, the total SSS SN counts are useful to constrain the poorly known CC SN rate evolution, by a combination with the relatively well determined SN Ia rate evolution. We have tested a variety of DTD models; some of them are based on the stellar evolution theory, and others are simple analytic functions often used in the literature. It is found that most of DTD models are consistent with the current data set, and hence we cannot set strong constraint on the type Ia SN progenitor. On the other hand, this rather week dependence on DTD models is an advantage when one tries to constrain CSFH parameters. It is required that $\\alpha$ (the SFR evolution index from $z=0$ to $\\sim 1$) is 3--4, with a considerable evolution of mean extinction of CC SNe, as $E(B-V) \\sim 0.2$ at local and $E(B-V) \\sim 0.5$ at $z \\sim 0.5$. Since we did not utilize any information from CSFH estimates by galaxy surveys, we can compare our result with those by galaxy surveys. Recent estimates based on UV luminosity are $\\alpha \\lesssim 2.5$, while those based on H$\\alpha$ or mid-infrared luminosity are close to our result, $\\alpha \\sim$ 3--4. These are nicely consistent with our finding of the significant evolution of extinction for CC SNe, indicating a strong evolution of extinction of star formation activity in the universe even at $z \\sim$ 0--0.5. The consistency between CSFH based on SFR in galaxies and SN rates is not trivial. Most indicators of galactic SFR trace the production of UV or ionizing photons and hence the formation of massive stars, which is the same as CC SNe. However, an evolution of IMF of massive stars could change the ratio of UV photon production to CC SN rate. Our result implies a roughly constant production efficiency of SNe per unit mass of star formation, and this would give some constraint on IMF evolution or metallicity effects. We have also demonstrated that, based on the counts of SN candidates without host galaxies, the contribution to the cosmic star formation activity from faint galaxies under detection limit or intergalactic star formation should not be significant. This is a clear advantage of CSFH constraint from supernova surveys, which cannot be obtained by CSFH studies based on galactic SFR estimates. The authors would like to thank L. Greggio for providing the data of delay time distribution and useful discussions. We also appreciate useful comments from A. Gal-Yam, D. Maoz, D. Poznanski, and an anonymous referee. A part of the SSS data was obtained as a part of the Supernova Cosmology Project (SCP), and we also thank the SCP members for useful discussions. This work was supported by the Grant-in-Aid for the 21st Century COE ``Center for Diversity and Universality in Physics'' from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan. T.O. has been supported by JSPS Research Fellowships for Young Scientists. T.T., N.Y., and M.D. are supported by the JSPS - USA bilateral programme." }, "0801/0801.3192_arXiv.txt": { "abstract": "{A good correlation has been found in star-forming galaxies, between the soft X-ray and the far infrared or radio luminosities. The soft X-ray emission in star-forming regions is driven by the heating of the diffuse interstellar medium, and by the mechanical energy released by stellar winds and supernova explosions, both directly linked to the strength of the star formation episode. } { We analyze the relation between the soft X-ray and far infrared luminosities as predicted by evolutionary population synthesis models, aiming first to test the validity of the soft X-ray luminosity as a star formation rate estimator, using the already known calibration of the FIR luminosity as a proxy, and second to propose a calibration based on the predictions of evolutionary synthesis models.} { We have computed the soft X-ray and far infrared luminosities expected for a massive starburst as a function of { evolutionary state, the efficiency of the conversion of mechanical energy into soft X-ray luminosity, the star formation history (instantaneous or extended) and dust abundance,}{} and we have compared these predictions with observational values for 62 star-forming galaxies taken from the literature. } {The observational \\lsxf\\ ratios are consistent with the model predictions under realistic assumptions (young starbursts, and efficiency in the re-processing of mechanical energy of a few percent), confirming the correlation between the diffuse soft X-ray emission and the star formation episode. } {The soft X-ray emission of the diffuse, extended gas surrounding massive star-forming regions, can be used as a star formation rate {tracer}. The empirical calibrations presented in the literature are supported by the predictions of evolutionary synthesis models, and by the analysis of a larger number of star-forming galaxies The calibrations are, however, biased towards galaxies dominated by relatively unevolved starbursts. } ", "introduction": "The onset of massive star formation episodes in galaxies drives their observational properties in almost any wavelength range. The UV and optical become dominated by the continuum of massive, hot and young stars, as well as by the presence of nebular emission lines. After a few Myr of evolution, red supergiant stars contribute to most of the near infrared emission. The heating of interstellar dust particles by the powerful UV photons induces the thermal re-emission of large amounts of energy in the { mid and far infrared} domain. The injection of ionizing photons into the surrounding gas generates thermal radio emission, which is replaced by non-thermal emission as the ionizing power of the burst declines and the more massive stars begin to explode as supernovae. The direct relation between the strength of the star formation episode and the intensity of the different observable parameters has allowed a number of star formation rate calibrators to be defined, such as UV continuum, emission lines intensity, far infrared or radio luminosities \\citep{Kennicutt98,Rosa02,Bell2003}. These calibrators have proven to be invaluable for statistical studies of the star formation history of the Universe. Star-forming regions are also the source of conspicuous X-ray emission, generated by individual stars, by the injection of large amounts of mechanical energy heating the interstellar medium, by supernova remnants, and by binary systems transferring mass to a compact primary \\citep{Cervino02,Persic02}. All of these individual components are in principle directly linked to the strength of the star formation episode, so that the X-ray luminosity could also be used as an estimator of star formation rates (\\sfr). \\begin{figure*} \\begin{center} \\includegraphics[width=8 cm,bb=5 39 786 521 dpi,clip=true]{8398fi1a.eps} \\hspace{1.0 cm} \\includegraphics[width=8 cm,bb=17 39 786 527 dpi,clip=true]{8398fi1b.eps} \\end{center} \\caption{Evolution of \\lfir\\ (left) and \\lsx\\ (right) for IB (dashed line, right axis) and EB (solid line, left axis) models. IB models predictions are shown normalized to $1$~\\msun\\ of gas transformed into stars, while for EB the luminosities are scaled to \\sfr\\ $= 1$ \\msun\\ yr$^{-1}$. \\lfir\\ has been plotted (from top to bottom) for \\ebv\\ $= 1.0$, $0.5$ and $0.1$. \\lsx\\ has been computed (from top to bottom) for \\xeff\\ $= 0.1$, $0.05$ and $0.01$. } \\label{lumin} \\end{figure*} Several authors have in recent years discussed the feasibility of using the X-ray luminosity as an \\sfr\\ estimator. {\\citet{Fabbiano02} already concluded from the analysis of 234 S0/a-Irr galaxies observed with {\\em Einstein}, that the correlation they found between the X-ray and the FIR luminosities in Sc-Irr galaxies was due to the young stellar populations in these objects.} \\citet{Ranalli03} proposed an empirical calibration of the soft ($0.5$--$2.0$ keV) and hard ($2$--$10$ keV) X-ray luminosities, based on their correlation with the far infrared (FIR) and radio luminosities, and using the known calibrations of these parameters as proxies. \\citet{Grimm03} studied the correlation between the number of high-mass X-ray binaries (HMXB) and the \\sfr, deriving different calibrations of the hard X-ray luminosity for low and high star formation rates. \\citet{Persic04} obtained a calibration of the hard X-ray luminosity as a \\sfr\\ estimator by assuming that most of the emission in this range is associated with HMXB, and scaling from the number of HMXB to the \\sfr\\ of our Galaxy. \\citet{Gilfanov2004} confirmed the {calibration} of the hard X-ray luminosity associated with HMXB, using slightly different slopes at high and low star formation rates. In a recent paper, \\citet{Persic07} found indeed that the collective hard X-ray emission of young point sources correlates linearly with the star formation rate derived from the far infrared luminosity. \\citet{Strickland04b} demonstrated that the luminosity of diffuse X-ray emission in star-forming galaxies is directly proportional to the rate of mechanical energy injection from the young, massive stars into the interstellar medium of the host galaxies. A similar result was found by \\citet{Grimes05} from the analysis of a sample of starburst galaxies of different types (from dwarf starbursts to ultraluminous infrared galaxies), which concluded that the mechanism producing the diffuse X-ray emission in the different types of starbursts was powered by the mechanical energy injected by stellar winds and supernovae into the surrounding medium. Recently \\citet{Rosa07} confirmed the reliability of the soft X-ray luminosity as an \\sfr\\ estimator from the analysis of a sample of star-forming galaxies in the {\\em Chandra Deep Field South} at redshifts $ z = 0.01 - 0.67$, using the UV continuum luminosity from {\\em GALEX} as a proxy. While hard X-ray emission from star-forming regions may be dominated by binary systems, diffuse soft X-ray emission is generated by reprocessed, mechanical energy from stellar winds and supernovae explosions. This mechanical energy is related to the strength of the burst of star-formation, and can be calculated using evolutionary population synthesis models. In this paper, we analyze the correlation between the soft X-ray and FIR luminosities in star-forming regions, both predicted by evolutionary synthesis models. We study the \\lsxf\\ ratio as a function of the {evolutionary state, the efficiency of the conversion of mechanical energy into soft X-ray luminosity, the star formation history (instantaneous or extended), and the dust abundance,} and compare the computed values with observations taken from the literature. { Our objective is to derive a calibration of \\lsx\\ as a tracer of the star formation rate, based on the predictions of evolutionary synthesis models, and to test the validity of the empirical calibration proposed by \\citet{Ranalli03}. } In Sect.~2 we describe the evolutionary synthesis models that we use in the present study, in Sect.~3 we present the observational data taken from the literature, and in Sect.~4 we discuss the predictions and the comparison with the observational values. {\\ Throughout this work we have assumed $H_{\\rm0} = 73$ km s$^{-1}$ Mpc$ ^{-1}$ to convert fluxes into luminosities.} \\begin{figure*} \\begin{center} \\includegraphics[width=8cm,bb=20 39 713 527 dpi,clip=true]{8398fi2a.eps} \\hspace{1.0 cm} \\includegraphics[width=8cm,bb=20 39 713 527 dpi,clip=true]{8398fi2b.eps} \\end{center} \\caption{Evolution of the \\lsxf\\ ratio computed for \\xeff\\ $= 0.01, 0.1$ and \\ebv\\ $= 0.1, 1.0$. Left: predictions for IB models; right: predictions for EB models. } \\label{ratio} \\end{figure*} ", "conclusions": "We have compared the \\lsxf\\ values measured in a sample of 62 star-forming galaxies with the predictions by our evolutionary synthesis models, aiming to analyze the validity of semiempirical and theoretical calibrations of \\lsx\\ as a star formation rate estimator. The main results can be summarized as follows: \\begin{enumerate} \\item The \\lsxf\\ ratios are strongly dependent on the efficiency in the conversion of the mechanical energy released by the young massive starburst into soft X-ray luminosity, by interaction of the stellar winds and supernova ejecta with the surrounding interstellar medium. From theoretical predictions and observational data, we expect an \\xeff\\ value of few percent in starburst galaxies. \\item \\lsxf\\ is also dependent on the evolutionary status of the star formation episode. It increases rapidly with time during the first $5$ Myr of evolution of a massive starburst, and shows a slower increase afterwards. After a (nearly) instantaneous burst of star formation, \\lsx\\ decreases slower than \\lfir, as long as there remains a population of massive stars able to collapse as supernovae (up to around 35~Myr). \\item When star formation proceeds at a nearly constant rate during extended periods of time, the \\lsxf\\ ratio is expected to stabilize after around $40$ Myr, when the number of massive stars that produce supernovae has reached an equilibrium between death and birth of new stars. { \\item The \\lsxf\\ values measured for the sample of star-forming galaxies are consistent with the predictions by the models under realistic conditions: relatively young and unevolved star formation episodes and \\xeff\\ values within $1$--$10$\\%. \\item A calibration of \\lsx\\ as a star formation rate estimator, based on the predictions of evolutionary synthesis models, has been derived. The calibration proposed by \\citet{Ranalli03} is consistent with the predictions for relatively unevolved, time-extended bursts of star formation. } \\end{enumerate}" }, "0801/0801.1474_arXiv.txt": { "abstract": "{} { We present the characteristics and some early scientific results of the first instrument at the Large Binocular Telescope (LBT), the Large Binocular Camera (LBC). Each LBT telescope unit will be equipped with similar prime focus cameras. The blue channel is optimized for imaging in the UV-B bands and the red channel for imaging in the VRIz bands. The corrected field-of-view of each camera is approximately 30 arcminutes in diameter, and the chip area is equivalent to a $23\\times 23$ arcmin$^2$ field. In this paper we also present the commissioning results of the blue channel. } { The scientific and technical performance of the blue channel was assessed by measurement of the astrometric distortion, flat fielding, ghosts, and photometric calibrations. These measurements were then used as input to a data reduction pipeline applied to science commissioning data. } { The measurements completed during commissioning show that the technical performance of the blue channel is in agreement with original expectations. Since the red camera is very similar to the blue one we expect similar performance from the commissioning that will be performed in the following months in binocular configuration. Using deep UV image, acquired during the commissioning of the blue camera, we derived faint UV galaxy-counts in a $\\sim 500$ sq. arcmin. sky area to U(Vega)$=26.5$. These galaxy counts imply that the blue camera is the most powerful UV imager presently available and in the near future in terms of depth and extent of the field-of-view. We emphasize the potential of the blue camera to increase the robustness of the UGR multicolour selection of Lyman break galaxies at redshift $z\\sim 3$. } {} ", "introduction": "The sensitivity of an optical system depends on a combination of the aperture and field-of-view (FoV). The imaging capabilities of existing or planned facilities are often limited by practical constraints. When the large collecting area of a telescope allows detection of faint sources, the field-of-view is typically less than $7\\times 7$ square arcminutes, and the UV sensitivity is low. Alternatively, wide-field imaging cameras onboard smaller telescopes are optimized to target brighter sources over a larger field-of-view (i.e. MegaCam at CFHT, \\cite{boulade}), and are unable to detect sources of faint magnitudes ($\\sim 28$) in particular in the UV. For these reasons an imager with a large FoV at an 8m class telescope is of fundamental importance to address the presently still open problems in stellar and extragalactic astronomy. The best example is the prime focus camera at the 8m Subaru telescope, Suprimecam (\\cite{subaru}). This imager is fast and has a FoV of $34\\times 27$ arcmin$^2$. Common science projects that have utilized this imager to date are the search of very high redshift galaxies, the study of the formation and evolution of galaxies, the investigation of the structure of the Universe, and the search for Kuiper Belt objects in the Solar system. The optical corrector cannot, however, simultaneously correct radiation of all wavelengths from UV to I-band. Due to this practical limitation, and to its low sensitivity in the blue band, Suprimecam does not provide imaging in the UV. At the end of the 1990s, it became clear that the binocular configuration of the Large Binocular Telescope (LBT) (\\cite{hill}), coupled with its mechanical design, provided a unique opportunity to justify a double prime focus camera capable of studying the widest-possible wavelength range from the UV down to the NIR H-band. The Large Binocular Camera (LBC, \\cite{ragazzoni,pedik,pedik04,ragazzoni06}) is a wide FoV instrument at the prime focus of the twin 8.4 meter Large Binocular Telescope (LBT). The LBT uses two 8.4-meter diameter honeycomb primary mirrors mounted side-by-side to produce a collecting area of 110 square meters equivalent to an 11.8-meter circular aperture. A unique feature of the LBT is that the light from the two primary mirrors can be combined optically in the center of the telescope to produce phased array imaging of an extended field. This requires minimal path length compensations, thus making interferometry easier than in completely independent telescopes. The requirement for an instrument such as LBC has been identified by several high-profile scientific programs that call for an increase in FoV and high-UV/IR sensitivity for deep imaging. These attributes are essential to programs studying a large FoV, to significant depth, over a wide spectral range, and can only be provided by an imager mounted at the prime focus of an 8m-class telescope. In Section 2 we provide a description of the two LBC cameras, while in Section 3 we detail the technical performance of LBC-Blue during commissioning observations in 2006. In Section 4 we analyze in detail the case for an UV deep imaging survey in an extragalactic field, and compare results with those obtained using different instruments and telescopes. We present our conclusions in Section 5. ", "conclusions": "In this paper we have described the first instrument at the LBT telescope, the prime focus large binocular camera (LBC). The instrument has a binocular configuration with two channels, the blue channel with a good overall efficiency in the UV band and the red channel with good efficiency in the V-z bands. We have also shown the technical characteristics of the blue channel derived from the commissioning data of LBC-Blue. \\begin{itemize} \\item The corrected optical field is $\\sim$ 30 arcmin of diameter with $<5$\\% loss of energy within $\\sim 25$ arcmin. The total throughput of the optical corrector is 84\\%. \\item The optical distortion is always $<1.75$\\% even at the edge of the field and it is removed with a specific SW package. \\item The optical quality ensures an energy concentration of 80\\% within a pixel of $\\simeq 0.2254$ arcsec in the overall corrected field. The active control of the optical quality can provide images as sharp as FWHM=0.5 arcsec even in the UV band. \\item Ghost images produced by the whole optical system are always negligible when glass filters are used (Bessel U,B,V). The total intensity of the primary brightest ghost is $\\sim 2.8$\\% when the wide interference UV filter is used. The primary ghost is centered on the original source position. Ghost images produced by the electronics cross-talk are as small as $3\\times 10^{ -5}$ and can be easily removed during data reduction. \\end{itemize} We have also described some scientific observations planned for the commissioning to assess the performance of LBC-Blue. \\begin{itemize} \\item We have obtained very deep UV galaxy counts in a deep pointing with a total exposure time of 3h reaching U(Vega)=26.5, after correction for incompleteness. The wide magnitude interval ($U(Vega)=19-26.5$) in our galaxy counts allowed a direct evaluation of the shape of the UV counts which shows a break in slope at about $U(Vega)=23.2$ with a change in slope from 0.62 to 0.22 at the faint end. \\item The same LBC area includes a quasar field where extensive study of Lyman break galaxies at redshift $z\\sim 3$ is available. We have reproduced with our UGR filter set the well known multicolour selection of LBGs showing that the robustness of the UV dropout method for the selection of star-forming galaxies at $z\\sim 3$ increases when very deep UV images can be used as obtained by LBC at an 8m class telescope like LBT. \\end{itemize}" }, "0801/0801.1368_arXiv.txt": { "abstract": "We have investigated the formation of close-in extrasolar giant planets through a coupling effect of mutual scattering, Kozai mechanism, and tidal circularization, by orbital integrations. Close-in gas giants would have been originally formed at several AU's beyond the ice lines in protoplanetary disks and migrated close to their host stars. Although type II migration due to planet-disk interactions may be a major channel for the migration, we show that this scattering process would also give a non-negligible contribution. We have carried out orbital integrations of three planets with Jupiter-mass, directly including the effect of tidal circularization. We have found that in about 30\\% runs close-in planets are formed, which is much higher than suggested by previous studies. Three-planet orbit crossing usually results in one or two planets ejection. The tidal circularization often occurs during the three-planet orbit crossing, but previous studies have monitored only the final stage after the ejection, significantly underestimating the formation probability. We have found that Kozai mechanism by outer planets is responsible for the formation of close-in planets. During the three-planet orbital crossing, the Kozai excitation is repeated and the eccentricity is often increased secularly to values close enough to unity for tidal circularization to transform the inner planet to a close-in planet. Since a moderate eccentricity can remain for the close-in planet, this mechanism may account for the observed close-in planets with moderate eccentricities and without nearby secondary planets. Since these planets also remain a broad range of orbital inclinations (even retrograde ones), the contribution of this process would be clarified by more observations of Rossiter-McLaughlin effects for transiting planets. ", "introduction": "\\label{sec:intro} More than 250 planets have been detected around both solar and non-solar type stars. Recent development of radial velocity techniques and accumulation of observations have revealed detailed orbital distribution of close-in planets. Figure \\ref{fig:exso}$a$ shows the distribution of semi-major axis and eccentricity for 236 planets \\footnote{http://exoplanet.eu/catalog-RV.php} discovered around solar type stars by the radial velocity techniques. The dotted line shows a pericenter distance $q=0.05$ AU. At $q \\la 0.05$ AU, many close-in planets have small eccentricities, which are accounted for by the circularization due to the tidal dissipation of energy within the planetary envelopes (Rasio \\& Ford 1996). The semi-major axis distribution of discovered extrasolar planets shows double peaks around 0.05 AU and 1 AU (e.g., Marcy et al. 2005; Jones et al. 2005). The location of the outer peak is determined by the observational limits, but the inner peak at around 0.05 AU, in other words shortage of planets between 0.06--0.8 AU, would be real. Since the close-in planets have relatively small eccentricities, the peak around 0.05 AU appears as a pile up at $\\log q/(1{\\rm AU})=-1.3$ also in pericenter distribution (Fig.~\\ref{fig:exso}$b$). The close-in planets are also found in the multi-planetary systems. Up to now, 23 multi-planet systems are reported with known planetary eccentricities (Fig.~\\ref{fig:multi}). The discovered close-in planets would have been formed at large distances beyond the ice line and migrated to shorter-period orbits (e.g., Ida \\& Lin 2004a, 2004b). A promising mechanism for the orbital migration is ``type II'' migration (e.g., Lin, Bodenheimer, \\& Richardson 1996). This is the migration with disk accretion due to gap opening around the planetary orbit caused by gravitational interaction between the planet and the protoplanetary gas disk. This mechanism can account for not only the pile-up of the close-in planets but also planetary pairs locked in mean-motion resonances. The migrating planets may have stalled as they enter the magnetospheric cavity in their nascent disks or interact tidally with their host stars. However, several close-in ($a \\la 0.1$AU) planets such as HAT-P-2, HD118203b, and HD162020b have moderate eccentricities ($\\ga 0.3$), but are not accompanied by nearby secondary large planets. HD17156b with $a \\simeq 0.15$AU may also belong to this group, since it has very large eccentricity ($\\simeq 0.67$) and $q$ as small as 0.05AU. It may be difficult for the planet-disk interaction alone to excite the eccentricities up to these levels. One possible mechanism for the excitation to the moderate eccentricity for close-in planets in multiple-planet systems is the passage of secular resonances during the epoch of disk depletion (Nagasawa, Tanaka, \\& Ida 2000; Nagasawa, Lin, \\& Ida 2003; Nagasawa \\& Lin 2005). Nagasawa and Lin (2005) showed that the relativity effect and the gravity of (distant) secondary planets excite the eccentricity of close-in planet orbiting around a slowly rotating host star. Although a main channel to form close-in planets may be type II migration, inward scatterings by other giant planets can be another channel, in particular for the close-in planets with moderate eccentricities. In relatively massive and/or metal-rich disks, multiple gas giants are likely to form (e.g., Kokubo \\& Ida 2002; Ida \\& Lin 2004a, 2004b, 2008). On longer timescales than formation timescales, the multiple gas giant systems can be orbitally unstable (Lin \\& Ida 1997; Marzari \\& Weidenschilling 2002). In many cases, a result of orbital crossing is that one of the planets is ejected and the other planets remain in well separated eccentric orbits (``Jumping Jupiters'' model; e.g., Rasio \\& Ford 1996; Weidenschilling \\& Marzari 1996; Lin \\& Ida 1997). Recent studies of the Jumping Jupiter model show pretty good coincidence with the observational eccentricity distribution (Marzari \\& Weidenschilling 2002; Chatterjee, Ford, \\& Rasio 2007; Ford \\& Rasio 2007). With the sufficiently small pericenter distance ($\\la 0.05$ AU), the planetary orbit can be circularized into the close-in orbit by tidal effects from the star (Rasio \\& Ford 1996). In this paper, the process of planet-planet scattering followed by the tidal circularization will be referred to as \"scattering model.\" To send a planet into the inner orbit, the other planets have to lose their orbital energy. Since inward scattering of a lighter planet is more energy saving, the systems which experienced the planet-planet scattering tend to have smaller planets inside. We might see the tendency in Figure \\ref{fig:multi}, but it is also true that radial velocity technique tends to detect heavier planets in outer regions. Note that since outer planets have negligible orbital energy, the energy conservation cannot restrict the semi-major axes of the outer planets and the outer planets may be located to the regions far beyond the radial velocity technique limit (Marzari \\& Weidenschilling 2002). Observation of Rossiter-McLaughlin effect for 5 transiting close-in planets including HAT-P-2 with eccentricity $e \\simeq 0.52$ show that their orbital planes are almost aligned with the stellar spin axes (Narita et al. 2007; Winn et al. 2005, 2006, 2007; Wolf et al. 2007), which may suggest that the scattering mechanism is not a main channel for formation of close-in planets, because the close scattering may excite orbital inclination as well as eccentricity (note that as shown in \\S 3.4, the orbital inclinations of close-in planets formed by the scattering model are not necessarily high). Since HD17156b with $e \\simeq 0.67$ is also transiting, the observation of Rossiter-McLaughlin effect will be important. More serious problem of the scattering model may be the probability for pericenter distances to become small enough (in other words, for eccentricities to become close enough to unity (e.g., $e \\ga 0.98$)) to allow tidal circularization. In a system with only two giant planets, the energy conservation law keeps the ratio of semi-major axes close to the initial value (see \\S \\ref{subsec:ppscat}). In such system, the probability for final planets with a large ratio of semi-major axes and very large eccentricity is $\\la 3$\\% (Ford, Havlickova, \\& Rasio 2001). In a system with more giant planets, the situation is slightly improved, especially when planets have non-zero inclinations. The close-in planets are very rare just after the scattering, but when they take into account the longer orbital evolution, the probability of the candidates for close-in planets increases to $\\sim$10 \\% (Weidenschilling \\& Marzari 1996; Marzari \\& Weidenschilling 2002; Chatterjee et al. 2007). The contributed planets are planets in Kozai state (see \\S 3.5). Here we revisit the scattering model through orbital integration of three giant planets with direct inclusion of the tidal circularization effects and analytical argument of Kozai effect. We find that the probability for formation of close-in planets is remarkably increased to $\\sim$30\\%. Although the previous studies were concerned with orbital states after stabilization by ejection of some planet(s), we find that the orbital circularization occurs through ``Kozai migration'' (Wu 2003; Wu \\& Murray 2003) caused by outer planets, mostly during three-planet orbital crossing, after a tentative separation of the inner planet and the outer planets. In the next section, we describe orbital instability of a planetary system (\\S \\ref{subsec:ppscat}) and the orbital evolution that becomes important after the system enters stable state: tidal circularization (\\S \\ref{subsec:tide}) and Kozai mechanism (\\S \\ref{subsec:kozai}). Following description of its models and assumptions, we present results of numerical simulations in \\S \\ref{sec:numerical}. We show typical outcome of the scattering (\\S \\ref{subsec:resultscat}), how the planets are circularized into short-period orbits (\\S \\ref{subsec:resulteach}), and its probability (\\S \\ref{subsec:resultcir}). In \\S \\ref{sec:analy} we present analytical arguments. We summarize our results in \\S \\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} ``Standard'' model for formation of close-in giant planets is that gas giants may be originally formed at several AU's beyond ice lines and migrate to the vicinity of the star. The most referred migration mechanism is type II migration. Here, we have investigated another channel to move planets to the stellar vicinity, ``scattering'' model (Rasio \\& Ford 1996), which is a coupled process of planet-planet scattering and tidal circularization. If orbital eccentricity is pumped up to values close to unity, because of the proximity of the pericenter the eccentricity and the semimajor axis of the planet are damped by the tidal effect from the star, almost keeping the pericenter distance to form a close-in planet with relatively small eccentricity. We newly found that Kozai mechanism is also one of key processes in the model. We have carried out orbital integration of three planets of Jupiter-mass, including the tidal damping in the integration. We follow the simulation setup by Marzari and Weidenschilling (2002). We found that in $\\sim$30\\% runs, close-in planets are formed, which is much higher probability than suggested by previous studies (Weidenschilling and Marzari 1997; Marzari and Weidenschilling 2002; Chatterjee et al. 2007), because in many cases, the circularization occurs during three-planet scattering, which was not monitored by previous studies. The three-planet scattering is usually terminated by ejection of one planet and the system enters a stable state. Previous studies were concerned only with the stable state. Since the tidal damping timescale is usually longer than that of orbital change in the chaotic stage of three-planet interaction, the tidal circularization does not occur during the chaotic stage, even if the pericenter distance happens to be small enough. But, we have found that when one planet is scattered inward, it often becomes separated from other outer planets. As long as the tidal dissipation is negligible, the isolation is only tentative, but its duration is long enough for the tidal circularization if the pericenter distance is small enough ($\\la 0.02$--0.04 AU for Jupiter-mass planets). We also found out that the isolated planet usually enters Kozai state. Even if the eccentricity of the planet is not excited enough by the close scattering that injected the planet to an inner orbit, the eccentricity can secularly increase to values close to unity during Kozai cycle, in particular, when the planet has an inclined orbit. Although the probability that the eccentricity takes values between 0.98 and 1.0 is quite low just after the scattering, that probability is significantly enhanced to 30\\% level during the Kozai cycle. Even if the Kozai cycle has relatively large angular momentum and the eccentricity does not take such high value, the outer planets eventually destroy the quasi-stable state and have the inner planet enter another quasi-stable state, in other words, another Kozai state. The new Kozai state can have relatively small angular momentum and raise the eccentricity high enough for the tidal circularization. The repeated Kozai mechanism enhances the probability for the tidal circularization to occur. Thereby, we have found much higher probability of formation of close-in planets than previously expected. Therefore, the scattering would contribute to formation of close-in planets, although a main channel may still be through type II migration. Non-negligible number of close-in planets without nearby secondary planets but with eccentricities $\\ga 0.3$ has been discovered. It is not easy for type II migration model to account for the high eccentricities, but the scattering model could account for them. As shown in sections 3.3 and 3.4, the tidal circularization can slow down by increase in $q$ before full circularization, leaving moderate eccentricities. The discovered close-in eccentric planets are generally massive (having masses larger than Jupiter-mass) among the close-in planets (Figure \\ref{fig:exso}a), so their circularization timescales are much shorter than those shown in Figure \\ref{fig:timescale}. Therefore, the results in our Sets T could be consistent with these discovered planets. Our results also suggest that the close-in planets have a broadly spread inclination distribution, including retrograde rotation. More observations of the Rossiter-McLaughlin effect for the transiting close-in planets, in particular for those with relatively high eccentricity such as HD17156b, will impose constraints on the contribution of the scattering model to close-in planets. The tidally dissipated energy is $\\sim G M_{\\ast} m/2a$ in total, where $a$ is final semi-major axis of the circularized close-in planet and $m$ is its planet mass. Since this energy is much larger than binding energy of the planet, it could be heat source for inflated close-in planets such as HD209458b, OGLE-TR-10, OGLE-TR-56, HAT-P-1b or WASP-1b, if the tidal circularization occurred relatively recently (orbital crossing followed by the circularization can start after long time after the formation of giant planets). Note that the tidal circularization model we used may not be sufficiently relevant. We did many assumptions for the usage of equations (\\ref{eq:ang}) and (\\ref{eq:eng}). The model is only valid for fully convective planets in highly eccentric orbits, since it consider only $\\ell=2$ f-mode and use impulse approximation. We apply the model even after the orbits are significantly circularized, which may overestimate effect of tidal dissipation. In the multiple encounters to the star, the energy and angular momentum can either increase or decrease (e.g., Press \\& Teukolsky 1997; Mardling 1995a, 1995b), but we include the tidal effect as it always gives decreasing effect. Because of these simplifications, our simulation may be a limiting case that the tidal force works most effectively, although inclusion of g-modes might strengthen the tidal effect. Our purpose of this paper is not a study of the detailed tidal damping process but to show another available channel to the close-in planets. We also need to take into account the relativity precession and J2 potential of the central star which prevent the Kozai mechanism, as well as long-term tidal disruption process (e.g., Gu, Lin, \\& Bodenheimer 2003). These issues are left for further studies." }, "0801/0801.0617.txt": { "abstract": "In 1996, a major radio flux-density outburst occured in the broad-line radio galaxy 3C\\,111. It was followed by a particularly bright plasma ejection associated with a superluminal jet component, which has shaped the parsec-scale structure of 3C\\,111 for almost a decade. Here, we present results from 18 epochs of Very Long Baseline Array (VLBA) observations conducted since 1995 as part of the VLBA 2\\,cm Survey and MOJAVE monitoring programs. This major event allows us to study a variety of processes associated with outbursts of radio-loud AGN in much greater detail than has been possible in other cases: the primary perturbation gives rise to the formation of a leading and a following component, which are interpreted as a forward and a backward-shock. Both components evolve in characteristically different ways and allow us to draw conclusions about the work flow of jet-production events; the expansion, acceleration and recollimation of the ejected jet plasma in an environment with steep pressure and density gradients are revealed; trailing components are formed in the wake of the primary perturbation possibly as a result of coupling to Kelvin-Helmholtz instability pinching modes from the interaction of the jet with the external medium. The interaction of the jet with its ambient medium is further described by the linear-polarization signature of jet components traveling along the jet and passing a region of steep pressure/density gradients. ", "introduction": "Direct evidence for the existence of bulk relativistic outflows along the jets in blazars and other radio-loud active galactic nuclei (AGN) comes from Very-Long-Baseline Interferometry (VLBI) observations. The first evidence for apparently superluminal structural changes was found from changes in the fringe visibility curves of 3C\\,279 and 3C\\,273 \\citep{Whi71,Coh71}. Subsequent higher-quality VLBI observations \\citep[see, e.g., compilation by ][and references therein]{Ver94} have established the ``core-jet'' type milliarcsecond-scale structure of compact extragalactic jets: the core being a bright and unresolved flat-spectrum component at the end of a linear structure, and the jet being composed out of individual steep-spectrum components or ``knots''. The knots frequently move away from the core with apparent velocities exceeding the speed of light. Monitoring observations of large source samples \\citep{Ver94,Jor01,Kel04,Pin07} have provided important statistical tools for probing relativistic beaming and the intrinsic properties of extragalactic radio jets \\citep{Coh07}, their intrinsic brightness temperatures \\citep{Hom06}, or their Lorentz factor distribution \\citep{Kel04} and luminosity function \\citep{Car07}. The relativistic-jet model \\citep[e.g.,][]{Bla79} has become the de-facto paradigm in multiwavelength research on blazars and other AGN, but VLBI observations have demonstrated that the basic concept of ballistically-moving isolated jet knots is clearly oversimplified: jet curvature \\citep[e.g.,][]{Ver94}, stationary components \\citep[e.g.,][]{Jor01}, and non-radial and accelerated motions \\citep[e.g.,][]{Kel04}, are found to be common features of relativistic jets. Within individual jets, there are often characteristic velocities suggesting the presence of an underlying continuous jet flow, but the ``components'' themselves most likely represent patterns moving at a different speed than the underlying flow, e.g., as hydrodynamically propagating shocks \\citep{Mar85,Hug85}. Recent years have brought major improvements in numerical simulations of relativistic jets \\citep[see, e.g.,][for a review]{Gom05}. It is now possible to simulate three-dimensional relativistic jets \\citep[e.g.,][]{Alo03} and to compute the relativistic processes \\citep[e.g.,][]{Gom97} that transfer hydrodynamic results into observed brightness distributions (e.g., relativistic light abberation and light travel time delays). In particular, interactions between strong perturbations or shocks with the underlying jet flow and the jet-ambient medium can be simulated \\citep{Agu01}. With these new techniques, it is now possible to compare the generation, propagation and evolution of emission features in simulated and observed relativistic jets. The nearby (z$=$0.049)\\footnote{Assuming $H_0=71$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_{\\rm M} = 0.3$, $\\Omega_\\Lambda = 0.7$ ($1{\\rm \\,mas} = 1.0{\\rm \\,pc}$).} broad-line radio galaxy 3C\\,111 (PKS B\\,0415+379) shows a classical FR\\,II morphology on kiloparsec-scales spanning more than 200$^{\\prime\\prime}$ with a highly collimated jet connecting the central core and the northeastern lobe in position angle 63$^\\circ$ while no counterjet is observed towards the southwestern lobe \\citep{Lin84}. This asymmetry is usually explained via relativistic boosting of the jet and de-boosting of the counter-jet. 3C\\,111 exhibits the brightest compact radio core at cm/mm wavelengths of all FR\\,II radio galaxies, a blazar-like spectral energy distribution \\citep{Sgu05}, and it was one of the first (and only) radio galaxies in which superluminal motion was detected \\citep{Goe87,Pre88}. Moreover, the (sub-) parsec scale jet of 3C 111 is intimately related to its high-energy emission: \\citet{Mar06} reports a disk-jet connection, similar to the well-established one in 3C 120 \\citep{Mar02}, in the sense that dips in the X-ray light curve indicate accretion events which are followed by VLBI jet component ejections. Recently, R.C.\\,Hartman \\& M. Kadler (in prep.) showed that the gamma-ray source 3EG\\,J0416+3650 can be decomposed into multiple individual sources inside the EGRET full-band point-spread function, revealing a significant signal from the nominal position of 3C\\,111 in the higher-resolution, high-energy band above 1\\,GeV. This association of 3C\\,111 with 3EG\\,J0416+3650, which had originally been suggested by \\citet{Har99} and \\citet{Sgu05}, makes 3C\\,111 one of the very rare radio galaxies detected at gamma-ray energies and supports the view that this source may be considered a lower-luminosity version of powerful radio-loud quasars. Here, we report the results from ten years of Very-Long-Baseline Interferometry (VLBI) observations of 3C\\,111 as part of the VLBA 2\\,cm Survey\\footnote{\\tt http://www.cv.nrao.edu/2cmsurvey/} \\citep{Kel98,Zen02,Kel04,Kov05} and its follow-up program MOJAVE\\footnote{\\tt http://www.physics.purdue.edu/MOJAVE} \\citep{Lis05,Hom06b}. We investigate the parsec-scale source structure during a major flux-density outburst and during its aftermath. We find that this outburst was associated with the formation of an exceptionally bright feature in the jet of 3C\\,111. A variety of processes (beyond the predictions of simple ballistic motion models) are observed and discussed in view of modern relativistic-jet simulations. In Sect.~\\ref{sect:obs}, our observations and the data reduction are described. A detailed report of the observational results is given in Sect.~\\ref{sect:results}. In Sect.~\\ref{sect:discussion}, we discuss the various processes observed in the jet of 3C\\,111 as a result of the outburst and during the propagation of the new jet feature along the jet. In Sect.~\\ref{sect:conclusions}, we put these results into the context of future simulations and observations with the goal of understanding the production mechanisms of AGN jets. ", "conclusions": "\\label{sect:conclusions} In this paper, we have investigated the parsec-scale jet kinematics and the interaction of the jet with its ambient medium in the broad-line radio galaxy 3C\\,111. Our analysis has demonstrated that a variety of processes influence the jet dynamics in this source: a plasma injection into the jet beam associated with a major flux-density outburst leads to the formation of multiple shocks that travel at different speeds downstream and interact with each other and with the ambient medium. The primary perturbation causes the formation of a forward and a backward shock (or rarefaction). The latter fades away so fast that is likely to remain undetected in minor ejections. A separate work by \\citet{Per07} focuses on the nature and characteristics of these initial components. Several parsecs downstream, the jet plasma enters a region of rapidly decreasing external pressure, expands into the jet ambient medium and accelerates. In the following, the plasma gets recollimated and trailing features are formed in the wake of the leading component. A particularly interesting aspect of the source 3C\\,111 in the light of this and other recent works is that it is one of the very rare non-blazar gamma-ray bright AGN. Besides Centaurus\\,A \\citep{Sre99} and the possible identification of NGC\\,6251 with the EGRET source 3EG\\,J1621+8203 \\citep{Muk02}, 3C\\,111 is the only AGN whose jet-system is inclined at a relatively large angle to the line of sight and that has a reliable EGRET identification: \\citet{Sgu05} reconsidered the possible identification of the EGRET source 3EG\\,J0416+3650 with 3C\\,111, which was first suggested by \\citet{Har99} but considered unlikely because of the poor positional coincidence. Very recently, R.\\,C.~Hartman \\& M.~Kadler (in prep.) found that 3EG\\,J0416+3650 is composed out of at least two distinct components. One of them is the dominant source above 1\\,GeV and is in excellent positional agreement with the location of 3C\\,111. Compared to blazars, the large inclination angle and the relatively small distance of 3C\\,111 allow us to resolve structures along the jet that are as small as parsecs in deprojection and which would be heavily blended with adjacent features in blazar jets. As demonstrated in this paper, VLBA observations of 3C\\,111 probe 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. From early 2008 on, the gamma-ray satellite GLAST \\citep{Lot07} is going to monitor the sky. If detected by GLAST, 3C\\,111 may become a key source in the quest for an understanding of the origin of gamma-rays from extragalactic jets. In addition, the combination of GLAST and VLBA data with spectral data at intermediate wavelengths (optical, IR, X-ray) may allow a better determination of jet parameters and relativistic beaming effects than in most blazars because of the higher linear resolution offered by this nearby and only weakly projected jet system. Our observations of 3C\\,111 are qualitatively in remarkable agreement with numerical relativistic hydrodynamic structural and emission simulations of jets such as the ones presented by \\cite{Agu01} and \\cite{Alo03}. Further progress is being made in the transition from two-dimensional to three-dimensional simulations of relativistic jets and in the development of new methods considering magnetic fields \\citep[e.g.,]{Lei05,Miz07,Roc08}, the equation of state for relativistic gases \\citep{Per07b}, and radiative processes \\citep[e.g.,][and Mimica et al. in preparation]{Mim04,Mim07}. But so far neither observational data nor simulations have reached an adequate level of detail and completeness in order to allow us a quantitative direct comparison of numerical models and observed relativistic jet structure and evolution. In particular, it is not feasible today to fit iteratively the parameters of relativistic magneto-hydro-dynamical (RMHD) jet simulations to match the brightness distribution observed for any individual source. The main reasons for this are a) the immense computational power required to conduct a realistic (i.e., sufficiently detailed) modern 3D jet simulation and b) the highly non-linear nature of RMHD plasmas and their evolution. Simulation results depend critically on the starting conditions like the exact velocity, composition, and transversal structure of the flow, the structure and strength of the magnetic field and the jet environment. Future development of computational power will allow us to use larger resolutions to decrease the numerical viscosities, and to implement nonlinear and microphysics processes into simulations. VLBA observations are capable of putting hard quantitative constraints on the input parameters for RMHD jet simulations if they are densely sampled over several years. Polarimetric observations at multiple radio frequencies may allow the effects of jet-intrinsic magnetic-field variations and external Faraday-screen inhomogenities or temporal variations to be disentangled. Such data at 15\\,GHz are on the way, e.g., as part of the next phase of the MOJAVE program, in which rapidly evolving sources like 3C\\,111 are being observed every two months." }, "0801/0801.1018_arXiv.txt": { "abstract": "{}{We study the effects of rotation on the outer convective zones of massive stars.}{We examine the effects of rotation on the thermal gradient and on the Solberg--Hoiland term by analytical developments and by numerical models.}{Writing the criterion for convection in rotating envelopes, we show that the effects of rotation on the thermal gradient are much larger and of opposite sign to the effect of the Solberg--Hoiland criterion. On the whole, rotation favors convection in stellar envelopes at the equator and to a smaller extent at the poles. In a rotating 20 M$_{\\odot}$ star at 94\\% of the critical angular velocity, there are two convective envelopes, with the bigger one having a thickness of 13.2\\% of the equatorial radius. In the non-rotating model, the corresponding convective zone has a thickness of only 4.6\\% of the radius. The occurrence of outer convection in massive stars has many consequences.}{} \\keywords {stars: evolution - convection - rotation } \\titlerunning{Convection in OB Stars} ", "introduction": "It is generally considered that the Cowling model applies to massive OB stars: i.e., a convective core surrounded by a large radiative envelope. However, long since stellar models have shown that massive stars have an outer convective envelope encompassing several percent of the stellar radius \\citep{Maeder80}. Also, \\citet{Langer97} has shown that an Eddington factor $\\Gamma ={\\kappa L}/{(4 \\pi c GM)}$ tending toward 1.0 implies convection. Our aim is to show that fast rotation amplifies the size of the convective envelope in OB stars as well as to develop anisotropic convective envelopes. Various limits can be considered about the effects of rotation and high luminosity on the stellar stability \\citep{Langer97,MMVI}: the $\\Gamma$--limit, which is the Eddington limit for $\\Gamma \\rightarrow 1$; the $\\Omega$--limit, which is reached by stars at rotational break--up with a small or negligible effect of the Eddington factor $\\Gamma$; the $\\Omega \\Gamma$-- limit, which applies to stars where both luminosity and rotation play significant roles. We show that not only the stars at the $\\Gamma$--limit, but also the stars at the $\\Omega \\Gamma$--limit and at the $\\Omega$--limit, have amplified external convective zones. The occurrence of outer convective envelopes in OB stars and their anisotropic structure lead to many astrophysical consequences: \\begin{itemize} \\item Convection generates acoustic modes that may allow asteroseismic observations of OB stars. \\item Convective motions may play a role in driving mass loss by stellar winds. \\item For stars close to the critical rotation, convective motions lower the effective break--up velocities. \\item An outer convective envelope may make a dynamo and contribute to some chromospheric activity generating an X--ray emission from OB--stars. \\item A convective envelope transports chemical elements and angular momentum. \\item The occurrence of outer convection may modify the von Zeipel theorem \\citep{vZ24}. \\end{itemize} A closer investigation is justified. We start by an analytical approach (Sect. \\ref{analytical}), and finish by two--dimensional models of 20 M$_{\\odot}$ rotating envelopes (Sect. \\ref{numerical}). ", "conclusions": "In stellar envelope of rotating stars, the effects of rotation on the thermal gradient arestronger and with the opposite sign with respect to the Solberg--Hoiland criterion, so that rotation favors convection instead of inhibiting it. The increase of the convective zone occurs mainly at the equator and also a bit at the poles. In a fast--rotating 20 M$_{\\odot}$ Pop I star, there are two equatorial zones covering a total of 16\\% of the stellar radius at the equator. There are several consequences of thees results to be examined in future. The outer convective motions may lower the escape velocity as well as the critical rotation velocity. The matter accelerated in the winds continuously goes through the convective zone in a dynamical process, suggesting that convection plays a role in accelerating the stellar winds and in producting the clumps in the winds. The convective pistons generate acoustic waves of periods of several hours to a few days. The density is very low, and it is thus likely that convection injects oscillations into the wind rather than into the interior." }, "0801/0801.4135_arXiv.txt": { "abstract": "The Bekenstein-Milgrom gravity theory with a modified Poisson equation is tested here for the existence of triaxial equilibrium solutions. Using the non-negative least square method, we show that self-consistent triaxial galaxies exist for baryonic models with a mild density cusp $\\rho \\sim {\\Sigma \\over r}$. Self-consistency is achieved for a wide range of central concentrations, $\\Sigma \\sim 10-1000\\mathrm{M_{\\odot}pc^{-2}}$, representing low-to-high surface brightness galaxies. Our results demonstrate for the first time that the orbit superposition technique is fruitful for constructing galaxy models beyond Newtonian gravity, and triaxial cuspy galaxies might exist without the help of Cold dark Matter. ", "introduction": "Constructing models of galaxies in triaxial equilibrium is a classical challenge in dynamics, either in Newtonian or Modified gravity. However, due to numerical constructions, few such models have been developed in Newtonian since the pioneering work of Schwarzchild(1979) on triaxial elliptical galaxies, and Zhao (1996) on the fast rotating triaxial bar of the Milky Way. Merritt \\& Fridman (1996) (hereafter MF96) considered a model with a central density cusp $r^{-1}$ and found that it is possible to construct a model of triaxial equilibrium galaxies self-consistently in Newtonian dynamics. Models with a steeper density cusp are not in equilibrium. This has generated interests in understanding cusp-triaxiality relation, the triaxiality-velocity anisotropy relation and the cusp-black hole relation. Recently, Capuzzo-Dolcetta et al. (2007) modeled $r^{-1}$ cuspy triaxial galaxies in $\\Lambda$ Cold Dark Matter (CDM) haloes. They took the model of MF96 as the luminous density distribution, and found that a model of cupsy triaxial galaxies with CDM halos is also self-consistent. A tough problem for $\\Lambda$CDM ($\\Lambda$ plus Cold Dark Matter) is that there is no physics to justify $\\Lambda$ the tiny cosmological constant or dark energy (see White 2007, Sarkar 2007). Although future particle physics experiments might well prove the existence of new species of particles and vacuum energy, their experimental-determined abundance could easily be factors of a few different from the 1:3 ratio as precisely required by fitting observations of Microwave background (Zhao 2006). This motivates exploring alternative theories like the co-variant version of Modified Newtonian Dynamics (MOND) in the mean-time. E.g., instead of more traditional interpretation of modified gravity (Bekenstein 2004), MOND was given the interpretation as a co-variant Dark Energy (DE) the ${\\bf V\\Lambda}$ model of Zhao (2007). This model uses General Relativity plus a non-uniform Dark Energy (DE) fluid (described by a four-vector flow) to give both the effects of DE and DM. Perturbations of such DE fluid bends space-time as an effective DM without actually invoking DM. Unlike $\\Lambda$CDM halos, which are positive and nearly round, the effective DM can sometimes be extremely flattened or negative in some regions of galaxies or clusters (Wu et al. 2007, 2008), although the data are not good enough to tell such negative regions yet (Nipoti et al. 2007a). Overall MOND and $\\Lambda$CDM halos are comparably successful in explaining the flat rotation curves of high surface brightness discs and satellites orbits at large radii (e.g., Angus et al. 2008). However, both theories are faced with severe challenges: $\\Lambda$CDM hinges on uncertain feedback on galactic scales, and even the maximum feedback falls short of explaining the velocity curves of low-surface brightness systems (Gnedin \\& Zhao 2002); Bekenstein's (2004) co-variant MOND seems to rely on non-relativistic neutrinos on large scale to match the weak lensing data (Angus et al. 2007) and the Cosmic Background Radiation (Skordis et al. 2006). \\footnote{although these problems appear resolvable in the ${\\bf V\\Lambda}$ model (Zhao 2007).} The tight correlation observed between the mass profiles of baryonic matter and dark matter at all radii in spiral galaxies (e.g., McGaugh et al. 2007; Famaey et al. 2007) is still the best case for the MOND paradigm of Milgrom (1983), which postulates that for accelerations below $a_0 = 1.2 \\times 10^{-10} {\\rm m} \\, {\\rm s}^{-2}$ the effective gravitational attraction approaches $(g_N a_0)^{1/2}$, where $g_N$ is the usual Newtonian gravitational field. Indeed, without resorting to galactic dark matter, this simple prescription is amazingly successful at reproducing galactic rotation curves over five decades in mass ranging from tiny dwarfs (e.g., Gentile et al. 2007) to early-type disk galaxies (e.g., Sanders \\& Noordermeer 2007) to massive ellipticals (Milgrom \\& Sanders 2003). Some outliers exist in models of gravitational lensing (Zhao et al. 2006, Chen \\& Zhao 2006, Shan et al. 2008). Lensing and structure formation are well-defined calculations in co-variant MOND (Halle \\& Zhao 2007). Galaxy formation simulations in classical MOND have been carried out too, and it is possible to form bars (Tiret \\& Combes 2007), and elliptical galaxies by mergers (Nipoti et al. 2007b), and spherical bulges by instability (Zhao, Xu \\& Dobbs 2008). Very little is known about the triaxial stationary equilibrium in non-Newtonian dynamics, although one speculates that such equilibrium might exist. In this paper, we test the existence of triaxial galaxies in the Bekenstein-Milgrom MOND theory, a well-defined classical theory with relativisitic and cosmological extentions. We apply the same density model given by MF96 with a fixed mass, and axis ratios. Our aim is to examine whether this model is self-consistent in MOND. The MOND theory is scale-dependent. When the gravitational acceleration is below $a_0$, the scaling deviates from the $r^{-2}$ Newtonian law. Here we consider models with different ratio of $a_0$ to the simulation unit G $\\times$ (Total Mass) / (Scale Radius)$^{2}$. We start with a nearly Newtonian system (high surface brightness) and finish with a deep-MOND system (low surface brightness). The rest of the paper is organized as follows. In \\S2, we present the density distribution of the galaxy and the corresponding potential and forces. The orbits are calculated in \\S3. The construction of self-consistent models is discussed in \\S4, and in \\S5, we give our conclusions and disscussions. ", "conclusions": "In practice, a parameter $\\delta$ is adopted to describe the departure form self-consistency, which is defined as in MF96: \\begin{equation} \\delta=\\frac{\\sqrt{\\chi^2}}{\\bar{M}}, \\end{equation} where $\\bar{M}$ is the average mass in each cell. Figure \\ref{delta} shows the departure from self-consistency as a function of the number of orbits. Note that there is a remarkably strong dependence on the number of orbits. The departure parameter $\\delta$ drops quickly when the number of orbits exceeds 4000. $\\delta$ is $~10^{-15}$ when the 6840 orbits are adopted in the optimization routine, which shows that all three triaxial models are self-consistent in MOND. Amazingly, we also find that the departure parameter $\\delta$ is insensitive to $a_0$. In other words, if one of the triaxial galaxies is self-consistent in MOND, then the same model with a different value of $a_0$ will also be self-consistent. Figure \\ref{ps_den} shows the contribution of mass for various orbital families in the self-consistent model. Chaotic orbits contribute a larger fraction of mass with the increase of energy. This shows a certain degree of consistency with that of Capuzzo-Dolcetta et al. (2007). However, discrepancies between our results and that of Capuzzo-Dolcetta et al. (2007) are apparent. There is a clear variational trend of the cumulative energy distribution of the various orbital families in Newton's case, however, no apparent variational trend for different orbits has been found in MOND. In Figure \\ref{sigma}, we show the radial profile of the rescaled radial velocity dispersion $\\sigma_r/V_{\\rm cir,\\infty}$(left panel) and the radial profile of the anisotropy parameter $\\beta=1-0.5()/$ (right panel), where $V_{\\rm cir,\\infty}$ is the circular velocity at infinity, and $V_r$ is radial velocity. The parameter $V^2$ is defined as $V^2=v_x^2+v_y^2+v_z^2$, where $v_x$, $v_y$ and $v_z$ are the three components of velocity. It can be seen that the profiles of the radial velocity dispersion are nearly flat in MOND, except in the central region. Recently, Capuzzo-Dolcetta et al. (2007) studied the same model as that used here, but in the Newtonian gravity. It can be seen that the profile of the velocity dispersion in our model shows a certain degree of consistency with that of Capuzzo-Dolcetta et al. (2007), but the discrepancies between our results and that of Capuzzo-Dolcetta et al. (2007) are apparent. There is central peak in our velocity dispersion profile. The global average of the velocity $V_{\\rm rms}$ is 170.49, 78.84 and 42.70 km/s for $a_0$=0.083, 0.833, and 8.333, respectively (see the last column of Table~\\ref{parameter}). The anisotropy parameter $\\beta\\geq0$ suggests that the stars have high radial velocity and the ratio of $/$ reaches a maximum value at $2-3$ kpc. Our anisotropy profiles are similar to that of Nipoti et al. (2007b) form dissapationless collapse in MOND (their figur 2) and in Newtonian gravity van Albada (1982) and Wang et al. (2008a). A gently rising beta is widely used (e.g., Milgrom \\& Sanders 2003). However, the peak in the anisotropy profile in our model is not seen in any simulations and observations to our best knowledge. This is might be indication of problem of MOND. Our models have a sudden change of the amount of box orbit near shells 10-12 (radius~2-3kpc), which could be the reason. In order to check the stability of our models, we used the Antonov's third law (Binney \\& Tremaine 1987). If the stellar system is stable, then its density $\\rho$ and potential $\\Phi$ satisfy everywhere the inequality ${\\rm d}^3\\rho_b/{\\rm d}\\Phi^3 < 0$. Figure~\\ref{drdp} shows that ${\\rm d}^3\\rho_b(0,y,0)/{\\rm d}|\\Phi(0,y,0)|^3$ is positive everywhere along the intermediate axis for our three models, since the values of potential are negative, our models are consistent with the law, which means the models appear globally stable. We will also be checking the stability of the models in collaboration with groups which have MOND N-body code. Effects due to pattern rotation and external field in MOND remain to be studied (Wang et al. 2008b). We have used the Schwarzschild approach to examine whether the triaxial galaxies are self-consistent in MOND. Using the Bologna Poisson solver to determine the potential and the accelerations at some discrete points, we used three-dimensional interpolation then obtain the potential and accelerations at arbitrary point. The orbits conserve energy to a very high level of accuracy, which shows that our method is feasible in calculating orbits using a grid-based MONDian potential. The departure parameter $\\delta$ shows that the triaxial galaxy models adopted in this paper in MOND are self-consistent. Our results show that it is theoretically allowed to have triaxial galaxies in rigorous equilibrium in a universe with either non-Newtonian gravity or certain models of Dark Energy (Zhao 2007)." }, "0801/0801.3653_arXiv.txt": { "abstract": "\\noindent Many models of baryogenesis rely on anomalous particle physics processes to give baryon number violation. By numerically evolving the electroweak equations on a lattice, we show that baryogenesis in these models creates helical cosmic magnetic fields. After a transitory period, electroweak dynamics is found to conserve the Chern-Simons number and the total electromagnetic helicity. We argue that baryogenesis could lead to magnetic fields of nano-Gauss strength today on astrophysical length scales. In addition to being astrophysically relevant, such helical magnetic fields can provide an independent probe of baryogenesis and CP violation in particle physics. ", "introduction": " ", "conclusions": "" }, "0801/0801.4559_arXiv.txt": { "abstract": "We present a survey of bright optical dropout sources in two deep, multiwavelength surveys comprising eleven widely-separated fields, aimed at constraining the galaxy luminosity function at $z\\approx7$ for sources at 5-10L*($z=6$). Our combined survey area is 225\\,arcmin$^2$ to a depth of $J_{AB}=24.2$ (3\\,$\\sigma$) and 135\\,arcmin$^2$ to $J=25.3$ (4\\,$\\sigma$). We find that infrared data longwards of 2$\\mu$m is essential for classifying optical dropout sources, and in particular for identifying cool Galactic star contaminants. Our limits on the number density of high redshift sources are consistent with current estimates of the Lyman break galaxy luminosity function at $z=6$. ", "introduction": "\\label{sec:intro} Direct observations of galaxies at very high redshift can cast light on the details of structure formation and early galaxy evolution. In recent years, the most commonly observed tracer of star formation at high redshift has been the Lyman Break galaxy (LBG) population \\citep{1999ApJ...519....1S}. These sources are selected photometrically based on a spectral feature interpreted as being the $\\lambda_\\mathrm{rest}=$1215.67\\AA\\ Lyman break. Neutral hydrogen along the line of sight leads to a forest of redshifted Lyman-$\\alpha$ absorption lines, blue continuum flux is suppressed, and a galaxy will exhibit red colours in a pair of filters bracketing the break. Depending on the relative depth of the imaging in different bands the galaxy will be seen to fall in magnitude between the red and bluewards bands, or even to drop below the detection limit. As a result, Lyman break galaxies are often termed `drops' or `dropouts' and their approximate redshift is determined by the band bluewards of the break. Hence `U'- and `G'-drops lie at $z\\approx3$ and $z\\approx4$ \\citep{1999ApJ...519....1S}, `V'- and `R'-drops at $z\\approx5$ \\citep{2003ApJ...593..630L,2004ApJ...600L.103G} and `I'-drops at $z\\approx6$ \\citep{2003MNRAS.342..439S,2006ApJ...653...53B}. The use of LBG samples to study star formation at ever-increasing redshifts has recently encountered a practical barrier. At $z\\approx7$ the Lyman break is redshifted to 1\\,$\\mu$m, a spectral region with poor sensitivity in both optical CCD detectors and near-infrared instruments. Photometry in the relatively broad photometric filters used longwards of 1\\,$\\mu$m effectively smears out the signatures of sharp-sided spectral features such as the Lyman break, making them difficult to distinguish from more gradual spectral features such as a Balmer break or molecular absorption in low redshift sources. The small field of view, high background, low sensitivity and low multiplexing of current near-infrared spectrographs further complicate the follow-up of Lyman break galaxy candidates. Sensitive, wide-format near-infrared imagers and multi-object spectrographs on the largest telescopes offer potential for $z=7$ Lyman Break galaxy searches, identifying bright near-infrared sources that drop in the optical bands (as discussed below in section \\ref{sec:col_sel}). Such surveys require knowledge of the $z=7$ luminosity function in order to estimate the required survey depth in a given area. High contamination rates in optical-dropout samples also complicate such surveys. While the most promising candidates will always require spectroscopic follow-up, it is clearly necessary to develop strategies for minimising the number of contaminating lower redshift sources from photometry alone. An important tool for fulfilling these requirements is the {\\em Spitzer Space Telescope}, which images at infrared wavelengths. However the 85cm aperture of {\\em Spitzer} and 1$''.$2$\\times$1$''.$2 arcsecond pixels of its IRAC instrument lead to blending and confusion issues at faint limits ($m_{AB}>24$). Existing deep field surveys incorporate sensitive imaging at wavelengths from 4000\\AA\\ to 8$\\mu$m (the four bands of the IRAC instrument). As a result, they can provide a first constraint on the surface density of bright galaxies at very high redshifts, and a test of the ability of {\\em Spitzer} to distinguish genuine $z>7$ sources from lower redshift contaminants. In this paper we present an analysis of two multiwavelength data-sets, identifying and classifying optical-dropout sources in eleven widely separated fields, averaging across cosmic variance. In section \\ref{sec:col_sel} we discuss the colour selection criteria which isolate high redshift galaxies. In sections \\ref{sec:ergs} and \\ref{sec:goods} we apply such criteria to the ESO Remote Galaxy Survey and the Great Observatories Origins Deep Survey respectively and consider the resulting candidate sources. Finally, in section \\ref{sec:discussion} we discuss the implications of this analysis for both the $z=7$ Lyman Break Galaxy luminosity function and future wide-area near-infrared survey. All magnitudes in this paper (optical and infrared) are quoted in the AB system \\citep{1983ApJ...266..713O}. We use a flat Universe with $\\Omega_{\\Lambda}=0.7$, $\\Omega_{M}=0.3$ and $H_{0}=70 h_{70} {\\rm km\\,s}^{-1}\\,{\\rm Mpc}^{-1}$. ", "conclusions": "\\label{sec:conc} The key points presented in this paper can be summarised as follows: i) We have identified an analysed a sample of $z$-band dropout sources in deep, multi-wavelength imaging, considering their suitability as potential $z>7$ galaxy candidates. ii) In an area of 360\\,arcmin$^2$ (the ERGS+GOODS surveys), we determine a strict upper limit of 1.4 potential $z=7$ galaxies and a likely limit of zero candidates to $J=24.2$. iii) In an area of 135\\,arcmin$^2$ (the GOODS survey), we determine a strict upper limit of 2.8 potential $z=7$ galaxies and a likely limit of no good candidates to $J=25.3$. iv) These results are consistent with current estimates of the luminosity function of Lyman break galaxies at $z=6$, for galaxies with $L>5L*$. v) The use of deep data longward of the $K$-band is extremely valuable if the number of high redshift candidates from deep surveys such as UKIDSS UDS from unmanageable to reasonable numbers for spectroscopic follow-up. \\subsection*{Acknowledgements} ERS gratefully acknowledges support from the UK Science and Technology Facilities Council (STFC). Based in part on observations made with the NASA/ESA Hubble Space Telescope, obtained from the Data Archive at the Space Telescope Science Institute. Also based in part on observations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory, CalTech. We thank the GOODS team for making these high quality datasets public. Research presented here has benefitted from the M, L, and T dwarf compendium housed at DwarfArchives.org and maintained by Chris Gelino, Davy Kirkpatrick, and Adam Burgasser. Results from the EDisCS/ERGS fields are based on observations made with ESO telescopes under programmes 166.A-0162 and 175.A-0706." }, "0801/0801.3186_arXiv.txt": { "abstract": "The unusual mushroom-shaped HI cloud, GW 123.4--1.5, is hundreds of parsecs in size but does not show any correlations to HI shells or chimney structures. To investigate the origin and velocity structure of GW 123.4--1.5, we perform three-dimensional hydrodynamical simulations of the collision of a high-velocity cloud with the Galactic disk. We also perform a parameter study of the density, radius, and incident angle of the impact cloud. The numerical experiments indicate that we reproduce the mushroom-shaped structure which resembles GW 123.4--1.5 in shape, size, position-velocity across the cap of the mushroom, and the density ratio between the mushroom and surrounding gas. GW 123.4--1.5 is expected to be formed by the almost head-on collision of a HVC with velocity $\\sim 100 \\kms$ and mass $\\sim 10^5 \\Msun$ about $5 \\times 10^7 \\yr$ ago. A mushroom-shaped structure like GW 123.4--1.5 must be infrequent on the Galactic plane, because the head-on collision which explains the mushroom structure seems rare for observed HVCs. HVC-disk collision explains not only the origin of the mushroom-shaped structure but also the formation of a variety of structures like shells, loops, and vertical structures in our Galaxy. ", "introduction": "GW 123.4--1.5 revealed by the Canadian Galactic Plane Survey \\citep{Tay03} is a mushroom-shaped HI structure composed of a stem and a cap and is dissimilar to any common shells or chimneys \\citep{eng2000}. Using the kinetic distance of the Galactic rotation model, \\citet{eng2000} derived properties of the mushroom cloud such as projected length, mass, velocity, mean excess column-density, and kinetic energy. The mushroom cloud extends a few hundred parsecs ($\\sim 350 \\pc$) at the assumed distance of $3.8 \\pm 1.2 \\kpc$ and has an HI mass of $\\sim 10^5 \\Msun$, a mean excess column-density of $N_{\\rm H} \\simeq 6 \\times 10^{20} {~{\\rm cm^{-2}}}$, and a kinetic energy of $\\sim 2 \\times 10^{50}$ ergs. As a model for the origin of the mushroom-shaped cloud GW 123.4--1.5, the gas buoyancy model \\citep[]{eng2000,dm2001} and the cloud collision model \\citep[hereafter KB2004]{kb2004} were proposed and numerical simulations were executed, respectively. \\citet{eng2000} carried out preliminary two-dimensional hydrodynamical simulations and showed that the origin of the mushroom-shaped GW 123.4--1.5 is the rise of buoyant gas from the explosion of a supernova above the Galactic midplane. Using three-dimensional hydrodynamical simulations of large-scale modeling of the interstellar gas in the galactic disk \\citep{da2000}, \\citet{dm2001} showed that the mushroom-shaped cloud can result from the buoyant rise of bubbles out of pools of hot gas created from a supernova or supernovae. However there is some difficulty in interpreting the origin of GW 123.4--1.5 by the buoyancy of a hot gas created by a single or multiple supernova explosions, because there are no observations of soft X-ray emission and IRAS sources near the base of GW 123.4--1.5. On the other hand, the collision of a high-velocity cloud (HVC) with the Galaxy, which was studied by \\citet{Ten86} for the first time, is not restricted to the regions of active star formation \\citep{Ten87}. KB2004 performed two-dimensional hydrodynamical simulations of the collision of a cloud with the Galactic disk and proposed that GW 123.4--1.5 is created by the impact of an intermediate-velocity cloud (IVC)\\footnotemark \\footnotetext{Historically, the clouds with $|v_{lsr}| \\geq 70 \\kms$ are classified into HVCs and those with $ < 70 \\kms$ are called intermediate-velocity clouds (IVCs).} with the Galactic disk. Also they showed that the velocity structure across the cap of the mushroom-shaped structure in their simulation is consistent with that of GW 123.4--1.5. In the buoyant models, the velocity characteristic of the cap is not obvious. Unlike the gas buoyancy model, cloud-disk collision can easily explain the formation of GW 123.4--1.5 without inference for X-ray emissions. In this study we extend the two-dimensional hydrodynamical simulation of KB2004 to three dimensions, in order to investigate the origin, dynamics, and velocity structure of a mushroom-shaped structure. Through numerical simulations and a parameter study, we can reproduce the size, shape, and position-velocity of the mushroom-shaped structure which resemble those of GW 123.4--1.5. Consequently, we can infer (1) how GW 123.4--1.5 was formed in the Galactic disk and (2) why a mushroom-shaped structure is so rare on the Galactic plane. In the next section we describe our models and numerical method. Simulation results are presented in \\S III, followed by a summary and discussion in \\S IV. ", "conclusions": "We perform three-dimensional hydrodynamical simulations for the impact of a HVC with the Galactic disk, in order to explore the formation of the mushroom-shaped HI cloud GW 123.4--1.5. The main results can be summarized as follows: \\begin{itemize} \\item A mushroom-shaped structure can be formed not only by a high-density IVC collision with the Galactic disk (KB2004) but also by a low-density HVC collision. \\item The resultant structures formed by cloud-disk collision are greatly affected by the density and incident angle of a colliding cloud. \\item GW 123.4--1.5 is expected to be formed by the almost head-on collision of a HVC with velocity $\\sim 100 \\kms$ and mass $\\sim 10^5 \\Msun$ about $5 \\times 10^7 \\yr$ ago. \\end{itemize} Recently, \\citet{as2005} have tried to find a mushroom-shaped HI structure like GW 123.4--1.5 in the Canadian Galactic Plane Survey. However, they did not find other mushroom shape structures. Our simulations could explain why mushroom-shaped structures are so infrequent in our Galaxy. A nearly head-on collision is needed to form a mushroom-shaped structure. In a rotating disk, the head-on collision of a cloud with the Galactic plane would be rare in a general way. HVC-disk collision can not only explain the origin of a mushroom-shaped structure but also the formation of a variety of structures like shells, loops and vertical structures in our Galaxy. For example, in model O30 (see the middle panel of Fig.\\,6), we can find an interesting structure, a loop which resembles that found by NANTEN in the central molecular zone within $\\sim 1 \\kpc$ from the Galactic center \\citep{Fuk2006}. The two loops of CO emission obtained with NANTEN have strong velocity gradients in a longitude-velocity diagram: $\\sim 80 \\kms$ per $250\\pc$ and $\\sim 60 \\kms$ per $150\\pc$ in each loop \\citep[see Fig.\\,2 of][]{Fuk2006}. Although the projected velocity distribution of our simulation ($\\sim 20 \\kms$) is somewhat smaller than that of observation ($\\sim 60 - 80 \\kms$), the size of the loop, $\\sim 300 (H_0/140\\,{\\pc}) \\pc$, is similar to that of the molecular loop. In order to explain the formation of the molecular loop, \\citet{Fuk2006} proposed a magnetic flotation model (due to magnetic buoyancy caused by the \\citet{Park66} instability). From our preliminary result of model O30, we also propose that HVC-disk collision is another possible model of formation of the molecular loop structure at the Galactic center. This issue will be studied in detail in an upcoming paper. The present study as well as the previous study (KB2004) are restricted to hydrodynamical simulations. When the magnetic field parallel to the Galactic disk exists \\citep[]{santi2000,franco2002,nozawa2005} , we also expect the magnetic buoyant effect (Parker instability). Both the impact of the cloud and the Parker instability \\citep[]{santi1999,santi04,santi07k} may create complicate structures in the Galactic disk such as worms, loops, shells, and chimneys. Moreover the interaction between the infalling HVCs and the magnetized disk must be an elementary process in the interstellar dynamics. Using high-resolution two-dimensional MHD simulation, \\citet{santi07} pointed out a possibility that the rain of compact HVCs onto the disk can maintain transonic turbulent motion in the warm phase of HI. In this sense, we believe that high-resolution three-dimensional MHD simulation will broaden our understanding of the structures formed by HVC collisions and bring us more interesting phenomena to be important in the interstellar dynamics. In the near future, we will perform three-dimensional magnetohydrodynamical simulations." }, "0801/0801.2978_arXiv.txt": { "abstract": "HD 179821 is an evolved star of unknown progenitor mass range (either post-Asymptotic Giant Branch or post-Red Supergiant) exhibiting a double peaked spectral energy distribution (SED) with a sharp rise from $\\sim8-20$ $\\mu$m. Such features have been associated with ejected dust shells or inwardly truncated circumstellar discs. In order to compare SEDs from both systems, we employ a spherically symmetric radiative transfer code and compare it to a radiative, inwardly truncated disc code. As a case study, we model the broad-band SED of HD 179821 using both codes. Shortward of 40 $\\mu$m, we find that both models produce equivalent fits to the data. However, longward of 40 $\\mu$m, the radial density distribution and corresponding broad range of disc temperatures produce excess emission above our spherically symmetric solutions and the observations. For HD 179821, our best fit consists of a $T_{eff}=7000$ K central source characterized by $\\tau_V\\sim1.95$ and surrounded by a radiatively driven, spherically symmetric dust shell. The extinction of the central source reddens the broad-band colours so that they resemble a $T_{eff}=5750$ K photosphere. We believe that HD 179821 contains a hotter central star than previously thought. Our results provide an initial step towards a technique to distinguish geometric differences from spectral modeling. ", "introduction": "HD 179821 (IRAS 19114+0002; V1427 Aql; SAO 124414; AFGL 2423) is an evolved star surrounded by gas and dust ejected during a phase of mass-loss. The luminosity of this object is undetermined as the Hipparcos parallax measurement ([0.18 $\\pm$ 1.12 mas]; \\citealt{1997A&A...323L..49P}) allows for any distance greater than 1 kpc. HD 179821 is either a close post-Asymptotic Giant Branch (post-AGB) star ($D=1$ kpc, $M_i\\sim3-4$ $M_\\odot$) or a distant, massive ($D=5-6$ kpc, $M_i\\sim20-30$ $M_\\odot$) post-Red Supergiant (post-RSG). The post-AGB phase of stellar evolution is a short ($\\sim10^3$ yrs) period in which initially intermediate-mass main sequence stars ($<8$ $M_\\odot$) transition from the Asymptotic Giant Branch to the planetary nebula (PN) phase. In contrast, the progenitors of Red Supergiants are massive main-sequence stars ($>8$ $M_\\odot$) evolving toward the Wolf-Rayet stage and may end their lives as supernovae. For a short review on the post-RSG status of HD 179821, see \\cite{2008arXiv0801.2315O}. In both phases, mass is ejected into the circumstellar environment. Distinguishing between these classes of objects can be difficult. In this paper, we model the broad-band SED of HD 179821 using two distinct radiative transfer codes. The first code, ``DUSTY\", computes radiative transfer through a spherically symmetric shell with canonical density profile $\\rho(r)\\propto r^{-2}$ (\\citealt{DUSTY}). The second code computes radiative transfer through an axisymmetric, flared disc with canonical density profile $\\rho(r)\\propto r^{-3/2}$ (\\citealt{2001ApJ...553..321D,2005ApJ...621..461D}). While we believe the mid-IR and radio imaging of HD 179821 indicates a roughly spherical nebula, the point of this paper is to determine whether spectral modeling, through the use of distinct radiative transfer codes, can constrain geometric features of evolved star nebulae. Thus, HD 179821 serves as a test case for this investigation. We model HD 179821 for both nebular geometries and investigate the degeneracies between the SED model fits. In addition, we present and compare the ISO spectrum of HD 179821 to the post-AGB object HD 161796 and post-RSG object IRC +10420. All three objects display a strong mid-IR excess and exhibit steep increases in their SED's beyond 8 microns. In Section 2, we discuss the enigmatic nature of HD 179821 in both the post-RSG and post-AGB scenarios. In Section 3, we describe our synthetic photospheres and extinction corrections. In Section 4, we present our SED model fits from the radiative transfer disc code. In Section 5, we present our model fits from DUSTY and identify wavelength regions in which degeneracies occur between the two codes. For intermediate mass stars below 8 $M_\\odot$, the late stages of stellar evolution are characterized by transitions from quasi-spherical mass-losing AGB stars to complex aspherical post-AGB objects. Many post-AGB objects feature toroidal density enhancements, bipolar jets and an array of other non-spherically symmetric features, the origin of which is unknown (\\citealt{2002ARA&A..40..439B}). In addition, most post-AGB objects display a large momentum excess above what would be supplied by radiation pressure (\\citealt{2001A&A...377..868B}). While the origin of the additional source of momentum is unclear, a binary companion is an attractive candidate as energy and angular momentum can transfer from the secondary to the primary (\\citealt{JN2006}). This hypothesis is supported by recent observational (\\citealt{2004ApJ...602L..93D,2007arXiv0709.1508D}) and theoretical (\\citealt{2006ApJ...650..916M,2006ApJ...645L..57S}) efforts suggesting that a significant fraction of PNe may be descendants of interacting binaries. Binary companions can influence mass loss in many ways. A common envelope phase, even for low-mass companions, can lead to equatorial outflows, poloidal outflows and the formation of discs (\\citealt{1999ApJ...524..952R}, \\citealt{JN2006}, \\citealt{JN2007}, \\citealt{2007arXiv0707.3792N}). For wider binaries, low-mass companions can induce spiral waves and convert amorphous dust to crystalline dust via annealing (\\citealt{2007arXiv0709.2292E}). For massive stars, the end stages of stellar evolution are poorly understood. The post-red supergiant (post-RSG) phase is among the most luminous and uncertain epochs of post-main sequence stellar evolution. Like their intermediate mass conterparts, post-RSG's are thought to be ejecting a substantial portion of their mass (\\citealt{2006A&A...456..549D}). However, a massive circumstellar envelope has only been detected (scattered light, IR emission, molecular line emission) around the post-RSG IRC +10420, thus making a comparison with HD 179821 difficult (\\citealt{1995ApJ...452..833K}). Mass-loss may be consistent with a radiation driven spherically symmetric outflow (\\citealt{2007A&A...465..457C}). However, if HD 179821 does indeed display a momentum excess in its ejected nebula (\\citealt{2001A&A...377..868B}), then it is not unreasonable for a binary companion to have influenced the outflow and produced asymmetries. Interestingly, the WFPC2 images of \\citealt{2000ApJ...528..861U} show collimated bipolar structures emerging from the dust shell. These may be remnant jets propagating through the envelope. There is also evidence for slight clumpy regions in both OH maser and mid-IR emission (\\citealt{2001MNRAS.328..301G,1995ApJ...452..833K}). A density asymmetry may also be present as the $^{13}$CO line profiles are asymmetric (\\citealt{1992A&A...257..701B}). The spectral energy distribution of HD 179821 exhibits a double-peaked shape, indicative of a stellar photospheric component and ejected dusty component. The spectral energy distribution (SED) is consistent with photospheric emission to $\\sim8$ $\\mu$m at which point there is a steep increase until the second peak at $\\sim$ 25 $\\mu$m. The overall SED of HD 179821 is thus remarkably similar to that of the transitional young stellar object CoKu Tau/4 (see Fig. 2 of \\citealt{2005ApJ...621..461D}), which has been well modeled as a dusty disc with an inner hole. The sharp, interior wall of the disc is illuminated by the central source and produces the excess infrared emission. HD 179821 also exhibits an evacuated interior region between the central object and hence, the system might be modeled similarly to CoKu Tau/4. However, the dust envelope of HD 179821 could also be a detached spherical shell and past modeling efforts have treated it as such (\\citealt{2002Ap&SS.281..751S,2007A&A...462..637B,2007A&A...465..457C}). ", "conclusions": "We have modeled the broad-band spectral energy distribution of HD 179821 using two distinct radiative transfer codes: one corresponding to a spherical shell geometry with $\\rho\\left(r\\right)\\propto r^{-2}$, and one to a disc geometry with $\\rho\\left(r\\right)\\propto r^{{-3}/{2}}$. Under these assumptions, both codes provided equally good fits for similar dust compositions and size distributions shortward of $40$ $\\mu$m. However, longward of $40$ $\\mu$m, only the spherical shell model reproduces the observations. The radial dust density profile and corresponding range of temperatures present in the disc provide excess emission above the spherical shell solution and cannot fit the data. A wavelength of $\\sim$ 40 $\\mu$m marks the boundary for discerning the model fits. It should be noted that the density profiles need not uniquely indicate a particular geometry. If the densities deviate from the anticipated theoretical values of $\\rho\\left(r\\right)\\propto r^{-2}$ (shell), $\\rho\\left(r\\right)\\propto r^{-3/2}$ (disc), then the infrared excess is likely coupled to the density distribution rather than the geometry. The effect of changing the density profile in both codes should be investigated before comparative spectral modeling is established as a reliable method for distinguishing geometric differences. Based on our detailed spectral modeling, we conclude that it is likely that the nebula around HD 179821 is a radiatively driven shell. The dust is mainly composed of small ($a_{max}=0.25$ $\\mu$m) amorphous, glassy silicates. Interior to the shell is an evacuated region in addition to the central radiation source. The central star is most likely a $T=7000$ K star obscured such that it appears as a $T=5750$ K star (optical depth of $\\tau_V\\sim1.95$ through the spherical shell). This may aid in explaining previous conflicting spectral classifications. The dimensions of our shell agree well with previous observations and provide a good fit to the ISO spectrum. We compared the ISO spectrum to that of HD 161796, a confirmed post-AGB object (e.g. Fig. \\ref{geometry}). In particular, the ISO spectra and broad-band SED look remarkably similar. Because our results scale with distance, our spectral modeling does not provide insight into whether HD 179821 is a post-AGB or post-RSG. However, the similarity of the ISO spectra suggests that HD 161796 and HD 179821 experienced a very similar mass-loss history." }, "0801/0801.0699_arXiv.txt": { "abstract": "An asymptotic treatment of thin accretion disks, introduced by Klu\\'zniak \\& Kita (2000) for a steady-state disk flow, is extended to a time-dependent problem. Transient growth of axisymmetric disturbances is analytically shown to occur on the global disk scale. The implications of this result on the theory of hydrodynamical thin accretion disks, as well as future prospects, are discussed. ", "introduction": "\\label{1sec} Thin accretion disks (AD) form whenever a sufficiently cool gas, endowed with a significant amount of angular momentum, is gravitationally attracted towards a relatively compact object. This situation is quite common in astrophysics and therefore the observational and theoretical study of accretion disks has been quite intensive. The necessary condition for accretion to take place is that angular momentum be extracted from the fluid swirling around the central object. To achieve accretion rates that are consistent with observations, the physical mechanism for such angular momentum transport must be more efficient (by many orders of magnitude) than just the one resulting from torques caused by microscopic viscosity. Already at the outset, when AD were theoretically proposed (Prendergast \\& Burbidge 1968, Pringle \\& Rees 1972), the enhanced effective viscosity was postulated to result from turbulence and was parametrized, using mixing length theory or a similar scheme, as detailed theoretical understanding of turbulence and the transition to it was lacking then (a situation that is still with us nowadays). The parametrization of the effective viscosity in disks in terms of a single parameter - $\\alpha$, introduced by Shakura \\& Sunyaev (1973), proved itself to be the most fruitful, giving rise to successful interpretations of many observational results (see Lin \\& Papaloizou 1996, Frank, King \\& Raine 2002, for reviews). In some cases (e.g. dwarf nova models), however, more complex (and cumbersome) prescriptions for the viscosity had to be employed. However, until the early 1990's the question what is the physical origin of turbulence (or, more precisely, the anomalous angular momentum transport) in AD has essentially remained unanswered. No definite linear instability has been identified for thin disks in which the swirling flow consist of Keplerian shear. The magneto-rotational instability (MRI), originally found (Velikhov 1959, Chandrasekhar 1960) for magnetic Taylor-Couette, (i.e., cylindrical) flows, has been shown by Balbus \\& Hawley (1991) to also operate in cylindrical AD, when the fluid is electrically conducting and for not too large initial magnetic fields. It has thus become the operating paradigm in the astrophysical community that purely hydrodynamical turbulence in Keplerian disks is altogether ruled out and since the problem of angular momentum transport in these objects relies on MRI driven magneto-hydrodynamical (MHD) turbulence, extensive numerical calculations are necessarily the main tool of research on this problem (see Balbus \\& Hawley 1998 and Balbus 2003, for reviews and references). Nonetheless, efforts to find a purely hydrodynamical transition to turbulent activity in thin AD have continued until the present time, more than a decade after Balbus , Hawley \\& Stone (1996) appeared to settle the matter. These have largely been motivated by the fact that a purely hydrodynamical AD flow has a `microscopic' Reynolds number ($\\rm Re$) of the order of $10^{14}$ or so, a quite amazing setting for a laminar flow. Among the ideas that have been put forward in this context some are based on the application to thin disk flows of the viewpoint, familiar to the fluid-dynamical community, that {\\em transient dynamics induced by perturbations}, i.e. strong transient growth (TG) in linearly stable shear flows may play an important r\\^ole in nonlinearly shaping the final dynamical state. TG is possible because the relevant linear operator governing the behavior of infinitesimal perturbations in these flows is non-normal and thus the usual modal approach essentially fails (see, e.g., Grossman 2000, Schmid \\& Hennigson 2001, and Criminale, Jackson \\& Joslin 2003) TG of disturbances has been discussed in the astrophysical literature within two quite different (but related) contexts: \\begin{description} \\item[{$\\bullet$}] Local disturbances experiencing large enough TG that can possibly trigger a subcritical nonlinear transition into turbulence in a linearly stable shear flow, i.e. via the so-called bypass transition (e.g., Chagelishvili {\\em et} al. 2003, Tevzadze {\\em et} al. 2003, Yecko 2004, Afshordi {\\em et} al. 2005, Mukhopadhyay {\\em et} al. 2005) \\item[{$\\bullet$}] Perturbations experiencing significant TG that can be excited either by some external agent, perhaps as secondary flows on a pre-existing 3D turbulence itself, however weak, may give rise to intense global dynamical activity (e.g. Ioannaou \\& Kakouris 2001, and see below). \\end{description} Other scenarios for inducing hydrodynamical activity in AD (some of them directly or indirectly related to TG as well) have also been proposed. Among them are those invoking baroclinic instabilities (e.g., Klahr \\& Bodenheimer 2003), strato-rotational instabilities (e.g., Dubrulle {\\em et} al. 2005, Umurhan 2006), formation and long sustenance of vortices and/or waves (e.g., Godon \\& Livio 1999, Bracco {\\em et} al. 1999, Li {\\em et} al. 2001, Umurhan \\& Regev 2004, , Barranco \\& Marcus 2005, Petersen, Stewart \\& Julien 2007, Bodo {\\em et} al. 2007, Lithwick 2007). Even though no definite and undisputed hydrodynamical mechanism, that can give rise to sufficient angular momentum transport in AD, has been identified so far, significant efforts along these lines are continuously being made by several groups of researchers. It appears that continued study of purely hydrodynamical processes in disks remains viable and worthwhile, in particular in view of the difficulties with MRI driven angular momentum transport in AD, which have recently been pointed out. Questions of insufficent numerical resolution in MHD disk simulations have been convincingly raised (Pesah, Chen \\& Psaltis 2007, Fromang \\& Papaloizou 2007). The decrease of transport with decreasing magnetic Prandtl number (${\\rm Pm}$) for various setups and boundary conditions and, in particular, the vanishing of this transport for ${\\rm Pm}\\ll 1$ (which is the case in many instances of astrophysical AD) has been demonstrated (Umurhan, Menou \\& Regev 2007, Lesur \\& Longaretti 2007, Umurhan, Regev \\& Menou 2007, Fromang {\\em et} al. 2007). Finally, serious doubts as to the viability of the local shearing box approximation (Goldreich \\& Lynden-Bell 1965) to the numerical study of accretion disk MHD turbulence (Coppi \\& Keyes 2003, King, Pringle \\& Livio 2007, Shu {\\em et} al. 2007, Regev \\& Umurhan 2007) have been raised. The purpose of this contribution is to draw attention to the possibility that a thin AD, in which a sub-critical transition to hydrodynamical turbulence occurs (for a sufficiently high ${\\rm Re}$) large transient amplification of global disturbances may induce recurrent or even persistent secondary flow activity. Lesur \\& Longaretti (2005) have explicitly demonstrated, using high resolution 3D numerical simulations, that such sub-critical transition to turbulence does appear in a model flow having Keplerian shear characteristics, but at a very high ${\\rm Re}$ (see also Rincon {\\em et} al. 2006). However, it appears that the efficiency of turbulent transport in these flows is insufficient for astrophysical purposes (i.e. AD). A similar conclusion can perhaps be drawn from the recent experimental study of Ji {\\em et} al. (2006). The very small value of the effective $\\alpha$ (measuring the angular momentum transport) in these flows is, as we shall show, the key necessary ingredient for the excitation of vigorous global (transient) secondary flows, atop the weak turbulent state. It is quite conceivable that the large (by orders of magnitude in the disturbance energy) TG may (nonlinearly) give rise to persistent dynamical activity, or at least be recurrently re-excited by external perturbations (see Ionnaou \\& Kakouris 2001, who advocated the latter possibility). It however remains to be shown, of course, that in the ultimate state angular momentum transport is appropriate for AD, i.e. the effective $\\alpha$ acquires a high enough value. The primary tool utilized here in order to facilitate an analytical treatment is the asymptotic expansion, where the dependent variables and governing equations are expanded in powers of a small quantity (here the measure of the disk's `thickness', $\\epsilon$ - see below). Exposing the resulting mathematical system (an initial value problem) to a set of initial conditions, in which the extreme geometry of the disk structure is taken into account (i.e. $\\epsilon \\ll 1$), we aimed at obtaining an analytically treatable problem. To achieve this goal we also assumed a polytropic relation between the pressure and density. The advantage of such an approach is obvious - the treatment can be essentially analytical and the responsible physical effects leading to any interesting dynamics may be transparently traced. The study reported on here is limited to axi-symmetric disturbances and the question of its ultimate development still remains open. We are now in the process of generalizing it to fully 3-D disturbances and it is clear that the present analytic work should be ultimately complemented with detailed and uncompromising 3-D numerical calculations with a proper treatment of energy generation and transfer. ", "conclusions": "\\label{4sec} We conclude with some remarks on possible improvements to the asymptotic analysis of the sort done here and prospects for the future, e.g., extensions to non-axisymmetric perturbations. Asymptotic expansions, when viable, are often very robust and provide a good approximation to the solution when truncation to only few leading terms is done. Obviously, when a term in the series becomes {\\em very} large it may `break its order', that is, become larger than a previous term and as such make the expansion invalid in this region. In our expansions successive terms ratios are of $\\order{\\epsilon^2}$, and thus the procedure's validity should not severely be limited even up to a growth factor of $\\sim 1000$ or so (in the velocity or density perturbations). This matter is further discussed in UNRS, here we just remark that the validity with respect to time for weakly viscous solutions are somewhat influenced by the thinness of the disk: smaller values of $\\epsilon$ mean that the solutions are valid for longer times after the initial disturbance. More importantly, however, for a given value of $\\epsilon$ one must not be too zealous or overreaching by attempting to infer the quantitative behavior of the disk for arbitrarily small values of $\\alpha$ - which, as one will recall, is formally assumed here to be an $\\order1$ quantity. Despite these caveats, the procedure when carried to higher order introduces corrections which are technically non-linear. Careful consideration must be undertaken in order to handle the response at these higher orders. This may entail treating the disturbance amplitudes for the lower order solutions (like $A(r)$ in $u_{1}'$) as {\\em weakly non-linear} governed by a second `slow' time (e.g. the amplitude is instead written as $A = A(r,\\tau)$ where $\\tau = \\epsilon^2 t$) in a manner analogous to the treatment of non-linear thick polytropes (e.g, Balmforth \\& Spiegel, 1996). The approach used here may be generalized in a number of directions. Allowing for non-axisymmetric perturbations, including the disk inner and outer boundary in some kind of boundary layer analysis and relaxing the polytropic assumption seem to be the most obvious generalizations. We have found the presence of prominent TG in the simplest cases. It is difficult to imagine that it will be suppressed in the more general conditions although the effect of radiative energy losses on TG must be carefully examined. The question concerning the ultimate fate of the transiently grown perturbations and their ability to induce a state of sustained complex dynamical activity in the disk remains open. In this context it is worthwhile to notice that since the TG decay times are of the order of hundreds of rotation periods, it is conceivable that AD, which are usually not isolated systems, may experience recurrent external perturbations on such time scales and in this way the dynamical activity may be sustained. Extensive numerical calculations of AD are however needed to decide if TG may lead, through non-linear processes, to sustained turbulence, or at least a dynamical state in which adequate angular momentum transport can be sustained. Such high-resolution global 3-D calculations are, however, still above the ability of the present computer power and it may be thus advantageous to also exploit sophisticated non-linear asymptotic methods to complement and guide them." }, "0801/0801.0520_arXiv.txt": { "abstract": "In this paper, we present a correlation between the spectral index distribution (SED) and the dimensionless accretion rate defined as $\\dot{m}={L_{bol}/L_{Edd}}$ for AGN. This quantity is used as a substitute of the physical accretion rate. We select 193 AGN with both broad H$\\alpha$ and broad H$\\beta$, and with absorption lines near MgI$\\lambda5175\\AA$ from SDSS DR4. We determine the spectral index and dimensionless accretion rate after correcting for both host galaxy contribution and internal reddening effects. A correlation is found between the optical spectral index and the dimensionless accretion rate for AGN, including low luminosity AGN ($L_{H\\alpha}<10^{41}{\\rm erg\\cdot s^{-1}}$ sometimes called \"dwarf AGN\" (Ho et al. 1997)). The existence of this correlation provides an independent method to estimate the central BH masses for all types of AGN. We also find that there is a different correlation between the spectral index and the BH masses for normal AGN and low luminosity AGN, which is perhaps due to the different accretion modes in these two types of nuclei. This in turn may lead to the different correlations between BH masses and optical continuum luminosity reported previously (Zhang et al. 2007a), which invalidates the application of the empirical relationship found by Kaspi et al. (2000, 2005) to low luminosity AGN in order to determine their BLR sizes. ", "introduction": "Because of the difficulty to calculate the physical accretion rate to the BH in an AGN, a dimensionless accretion rate can be defined and estimated based on the bolometric luminosity and BH mass, $\\dot{m} = \\frac{L_{bol}}{L_{Edd}}$. This parameter can be used as the substitute for the actual accretion rate. Moreover, the dimensionless accretion rate is an important parameter in the scheme of the so called unified model for AGN (Antonucci 1993, Quintilio \\& Viegas 1997, Urry \\& Padovani 1995), and also the 4D Eigenvector 1 Scheme (e.g., Dultzin-Hacyan et al. (2007). The difference in accretion rate leads to the principal difference between high luminosity QSOs and low luminosity Seyfert galaxies. It seems to be also one of the main physical parameters underlying the 4D Eigenvector 1 Scheme as explained recently in Dultzin-Hacyan et al. (2007). In order to obtain the dimensionless accretion rate $\\dot{m}$, two other parameters must be calculated first: BH mass $M_{BH}$ and the bolometric luminosity $L_{bol}$. The common method to estimate the bolometric luminosity $L_{bol}$ of AGN is based on the continuum luminosity from the nuclei: $L_{bol}\\sim 9\\times L_{5100\\AA}$ given by Kaspi et al. (2000) and confirmed by Shang et al. (2005) for QSOs. Recently, the relation was used by Bonning et al. (2007) to study the correlation between the accretion disk temperatures and the continuum colors in QSOs. However, it should be stressed that the relation does not hold for ALL kinds of AGN, in particular, for low luminosity AGN (Ho et al. 1997a, 1997b, Ho 1999), because of the different Spectral Energy Distribution (SED) (the lack of the Big Blue Bump). Several methods are used to estimate BH masses of AGN. The most reliable method is based on the stellar velocity dispersion of the bulge of the host galaxy first presented by Ferrarese \\& Merritt (2000) and Gebhardt et al. (2000), then confirmed by Tremaine et al. (2002) and Merritt \\& Ferrarese (2001) etc. \\begin{equation} M_{BH} = 10^{8.13\\pm0.06}(\\frac{\\sigma}{200{\\rm km\\cdot s^{-1}}})^{4.02\\pm0.32} {\\rm M_{\\odot}} \\end{equation} which indicates a strong correlation between BH masses and bulge masses (H\\\"{a}ing \\& Rix 2004, Marconi \\& Hunt 2003, McLure \\& Dunlop 2002, Laor 2001, Kormendy 2001, Wandel 1999) etc. However we should note that the relation of $M_{BH} - \\sigma$ is obtained through the results of nearby inactive galaxies. Whether the relation can be applied to far away active galaxies is an interesting question. So far, there are a few dynamical mass estimates of central black holes of broad line AGN, and the BH masses are consistent with the BH masses estimated from the relation of $M_{BH} - \\sigma$, although the uncertainties are still large. In addition, we should say that the objects in our sample described in the following section are not high luminosity and high redshift QSOs, thus the correlation between central black hole and bulge of host galaxy can be reasonably considered to hold. The other methods are based on the assumption of virialization of the Broad Line Emitting Regions (BLRs), or at least part of them (Peterson et al. 2004, Onken et al. 2004, Sulentic et al. 2006, Dultzin-Hacyan et al. 2007). In order to calculate the parameter of dimensionless accretion rate, the BH mass is necessary. However, for high luminosity and high redshift AGN, it is difficult to measure the stellar velocity dispersions. The assumption of virialization is applied for QSOs. In order to estimate BH masses of QSOs based on the assumption of virialization, the most convenient way is to use the equation: \\begin{equation} \\begin{split} M_{BH} &= f\\times\\frac{R_{BLRs}\\times\\sigma_{b}^2}{G} \\\\ &= 2.15\\times10^8(\\frac{\\sigma_b}{3000{\\rm km\\cdot s^{-1}}})^2(\\frac{L_{5100\\AA}}{10^{44}{\\rm erg\\cdot s^{-1}}})^{0.69} {\\rm M_{\\odot}} \\end{split} \\end{equation} There are, however, some caveats with this method. First, the question of whether the relation $R_{BLRs}\\sim L_{5100\\AA}^{0.69}$ found by Kaspi et al. (2000, 2005) can be applied for all AGN, in particular high redshift ones. In an attempt to answer this question, we have found that the relation is not valid for some special kinds of AGN, such as the low luminosity AGN (Zhang, Dultzin-hacyan \\& Wang 2007a, Wang \\& Zhang 2003) and the AGN with double-peaked low ionization emission lines (Zhang, Dultzin-Hacyan \\& Wang 2007b). Second, the estimation of the BH masses of high redshift AGN by means of Equation (2) will lead to BH masses larger than $10^{10}{\\rm M{\\odot}}$ (meaning $\\sigma>600{\\rm km\\cdot s^{-1}}$), which leads to unreasonable masses of the bulge larger than $10^{13}{\\rm M_{\\odot}}$ (Netzer 2003, Sulentic et al. 2006). For this reason, finding another parameter which can be observationally determined, related to the dimensionless accretion rate is an important task and is the main objective of this paper. The accretion rate is determined by two properties: the continuum luminosity $L_{5100\\AA}$ and the BH mass $M_{BH}$. The continuum luminosity can be calculated from the observed spectra as discussed in the the next section. Thus, in order to obtain a reliable result, we select Equation (1) to estimate the central BH masses of AGN rather than Equation (2). The accretion disk model has been widely accepted as the standard model for AGN. In the NLTE (Non Local Thermodynamic Equilibrium) accretion disk mode, the generated SED (Spectral Energy Distribution) is based on three main parameters: BH masses $M_{BH}$, accretion rate $\\dot{M}$ and the viscosity parameter $\\alpha$. An expected result is that there should be a correlation between the spectral index and the accretion rate $\\dot{m}$. In this paper, we answer the question whether the observed spectral index can be used to trace the dimensionless accretion rate. In section II, we present the data sample. Section III gives the results. Finally the discussion and conclusions are given in Section IV. In this paper, the cosmological parameters $H_{0}=70{\\rm km\\cdot s}^{-1}{\\rm Mpc}^{-1}$, $\\Omega_{\\Lambda}=0.7$ and $\\Omega_{m}=0.3$ have been adopted. ", "conclusions": "There are 38 low luminosity AGN with $L_{H\\alpha}<10^{41}{\\rm erg\\cdot s^{-1}}$ (Ho et al. 1997a, 1997b and Ho 1999), which are shown in solid circles in Figure 2. From the figure, we can see that there is no difference in the correlation between spectral index and $\\frac{9\\times L_{5100\\AA}}{L_{Edd}}$ for normal AGN and low luminosity AGN. If the bolometric luminosity of low luminosity AGN was different from $9\\times L_{5100\\AA}$, all the low luminosity AGN would deviate from the correlation for normal AGN, due (probably) to different accretion modes. Even if there is a different accretion mode for low luminosity AGN (as suggested by the lack of the big blue bump in the spectra of low luminosity AGN as shown in Ho (1999)), the bolometric luminosity of low luminosity AGN can also be calculated using $L_{bol}\\sim k\\times L_{5100\\AA}$. Otherwise, we could not find the same correlation between spectral index and $\\frac{9\\times L_{5100\\AA}}{L_{Edd}}$ for low luminosity AGN and normal AGN. According to the accretion disk model, the output SED is the result of the convolution of other parameters as well, such as the central BH mass, the viscosity in the disk, and the inclination angle. However there is no correlation between the spectral index and central BH masses, which is shown in Figure 5. The Spearman Rank Correlation Coefficient is less than 0.1 with $P_{null}>60\\%$ for all objects in our sample. An interesting result is that there is actually a negative trend (anticorrelation) between BH masses and the spectral indexes for the 38 low luminosity AGN. The coefficient is about -0.54 with $P_{null}\\sim4.97\\times10^{-4}$, -0.52 with $P_{null}\\sim8.62\\times10^{-4}$ and -0.55 with $P_{null}\\sim3.98\\times10^{-4}$ for $\\frac{F_{5100\\AA}}{F_{6800\\AA}}$, $\\frac{F_{4400\\AA}}{F_{5100\\AA}}$ and $\\frac{F_{4400\\AA}}{F_{6800\\AA}}$ respectively. Because of the positive correlation between the spectral index and the accretion rate, a negative correlation between the spectral index and central BH masses could be expected for all AGN. However our results indicate that this expectation is only valid for low luminosity AGN. The reason is probably related to the correlation between BH masses and continuum luminosity. For Normal AGN, there is strong correlation between the BH masses and the continuum luminosity (Peterson et al. 2004). However for low luminosity AGN, this correlation is much weaker (Zhang, Dultzin-Hacyan \\& Wang 2007a). The correlation between the central BH masses and the internal continuum luminosity is shown in Figure 6. The coefficient is about 0.47 with $P_{null}\\sim9.05\\times10^{-10}$, however, the coefficient is only 0.12 with $P_{null}\\sim49\\%$ for the 38 low luminosity AGN. The same result for low luminosity AGN can be found in Panessa et al. (2006). In their paper, they selected all the low luminosity Seyfert galaxies from Ho, Filippenko \\& Sargent (1997a, 1997b), and found that there is NO correlation between the X-ray or optical emission line luminosities (especially [OIII]$\\lambda5007\\AA$ line) and BH masses. The result also confirms that there is a different accretion mode for normal and low luminosity AGN. To estimate the effects of the inclination angle of the accretion disk is difficult. However under the assumption that the narrow line emission region is isotropic, we can check the correlation between the continuum luminosity and the luminosity of narrow emission lines. If the objects have very different inclination angles of the accretion disk, a loose correlation between the continuum luminosity and the luminosity of narrow emission line should be expected. Here we show the correlation between $L_{5100\\AA}$ and the luminosity of narrow H$\\alpha$ in Figure 7. Although it is more common to use the [OIII]$\\lambda5007\\AA$ as an isotropic estimator of AGN luminosity, we prefer to use the narrow component of H$\\alpha$ because of the following reason. [OIII] emission line frequently cannot be fitted by one single gaussian function, because it has extended wings (Greene \\& Ho 2005a). The two components are not emitted from the same region. The extended component is probably emitted from the far-side of the BLRs. Thus when we fit the [OIII] line, two gaussian functions are applied as described in Section II. We thus select the narrow component of H$\\alpha$ rather than [OIII] to test the effects of inclination angle. A strong correlation can be confirmed. The spearman Rank Correlation Coefficient is about 0.89 with $P_{null}\\sim0$ for normal AGN, and about 0.67 with $P_{null}\\sim4.04\\times10^{-26}$ for the 38 low luminosity AGN. The best fit to the correlation (after considering the error in the determination of the luminosity of narrow H$\\alpha$) is given by: \\begin{equation} \\log(L_{H\\alpha_N}) = (1.373\\pm0.032) + (0.915\\pm0.003)\\times\\log{L_{5100\\AA}} {\\rm erg\\cdot s^{-1}} \\end{equation} This result indicates that the effects of the inclination angle can be neglected for the correlation between the spectral index and the dimensionless accretion rate. The correlation between spectral index and dimensionless accretion rate found in this research provides another independent method to estimate the central BH masses of AGN. The spectral index and continuum luminosity can be directly determined from the observed spectrum in the optical band, and subsequently the Eddington Luminosity, i.e. BH masses, can be determined by means of the correlation we found. This method has the advantage of being independent of the different correlations between the size of the BLRs and the continuum luminosity in Equation (2). Also, this method can be applied when it is not possible to measure the stellar velocity dispersion of the bulge. In future work, we will estimate the BH masses of QSOs with higher redshift using this method to solve the problem of why virial BH masses of QSOs estimated by Equation (2) lead to BH masses larger than $10^{10} {M_{\\odot}}$, while observational results (fortunately) seem to contradict this result (Dultzin-Hacyan et al. 2007). Finally, a simple summary is as follows. We first select 193 AGN with both broad H$\\alpha$ and broad H$\\beta$, and with apparent absorption MgI$\\lambda5175\\AA$ from SDSS DR4. Then after the determination of the spectral index (after the correction of internal reddening effects through the Balmer decrements for broad Balmer emission lines, and after the subtraction of stellar component) and dimensionless accretion rate ($\\frac{9\\times L_{5100\\AA}}{L_{Edd}}$), we find a strong correlation between these parameters for AGN, which provides another independent and method to estimate the central BH masses of AGN." }, "0801/0801.2713_arXiv.txt": { "abstract": "{New candidate variable stars have been identified in the Small Magellanic Cloud cluster NGC121, by applying both the image subtraction technique (ISIS, Alard 2000) and the Welch \\& Stetson (1993) detection method to HST WFPC2 archive and ACS proprietary images of the cluster. The new candidate variable stars are located from the cluster's Main Sequence up to Red Giant Branch. Twenty-seven of them fall on the cluster Horizontal Branch and are very likely RR Lyrae stars. They include the few RR Lyrae stars already discussed by Walker \\& Mack (1988). We also detected 20 Dwarf Cepheid candidates in the central region of NGC121. Our results confirm the ``true\" globular cluster nature of NGC121, a cluster that is at the young end of the Galactic globulars' age range. ", "introduction": "NGC121 is a key cluster to understand the Star Formation History (SFH) occurred in the SMC through the study of the similarities and differences with respect to the Milky Way clusters. In fact, with an estimated age of about 10 Gyrs, it locates at the transition between open and globular clusters in the MW. Moreover, as emphasized by Stryker et al. (1987), NGC121 very likely marks the boundary between ``old'' Population II (containing RR Lyrae but not carbon stars) and intermediate-age populations (containing carbon stars but not RR Lyrae stars). Up to now, only four RR Lyrae stars with light curves derived from ground-based observations, were known in NGC121 (Walker \\& Mack 1988), and references therein). As part of a coordinated HST and ground-based effort, we have observed the SMC with the ACS (Cycle13, GO prog.10396, PI Gallagher), pointing at 7 star clusters of different age and metallicity (including NGC121) and 7 fields in various galactic locations (center, periphery, wing and bridge towards the LMC), to derive their key evolution parameters and SF histories (see e.g. Glatt et al. 2007, submitted). The ACS observations of NGC121 were taken in time-series fashion in order to use them, along with the existing WFPC2 archive data of the cluster, to detect new variable stars in the cluster. ", "conclusions": "\\begin{figure} \\includegraphics[width=6cm]{fig2_fiorentino_poster.ps} \\caption[]{Light curves of fundamental-mode RR Lyrae stars. {\\it Upper panel}: new RR Lyrae star with $P \\sim$ 0.63 days. {\\it Lower panel}: the fundamental-mode RR Lyrae star V37 discovered by Walker \\& Mack (1988).} \\label{map} \\end{figure} The NGC121 candidate variable stars were identified by applying both the image subtraction technique (ISIS, Alard 2000) and the Welch \\& Stetson (1993) detection method. We identified about 50 candidate variables in NGC121. Period search was performed using GRaTiS (Graphical Analyzer of Time Series) a private software developed at the Bologna Observatory (see e.g. Clementini et al. 2000). Twenty-seven of the candidate variables are located on the cluster Horizontal Branch, thus are very likely RR Lyrae stars. They include the few RR Lyrae stars already discussed by Walker \\& Mack (1988). We also detected 20 Dwarf Cepheid candidates in the central region of NGC121, and recovered 34 Blue Straggler star candidates found by Shara et al. (1998). DCs are late-A and early-F type stars, that populate the instability strip near or slightly above the zero-age main sequence and have magnitude from 0.2 to about 3 mag fainter than the RR Lyrae stars. In the CMD they fall where the instability strip crosses the region of the Blue Straggler stars. These variables have typical periods in the range from 1 to 6 hours, thus our data sampling allow to sample their light curves better than for the RR Lyrae stars. \\begin{figure} \\includegraphics[width=6cm]{fig3_fiorentino_poster.ps} \\caption[]{Light curves of two Dwarf Cepheids. {\\it Upper panel}: SX Phoenicis star with $P$=0.05 days. {\\it Lower panel}: $\\delta$ Scuti star with $P$=0.19 days. } \\end{figure} DCs are divided into: 1) $\\delta$ Scuti stars, which are metal rich, young, Population I stars with typical periods longer than 0.1 day usually observed in open clusters and in the field of the MW; and 2) SX Phoenicis stars, which are metal poor, Population II variables, generally observed in GCs, with typical periods in the range from 0.03 to 0.09 days (Poretti et al. 2006). NGC121 seems to host both types of DCs. Fig. 1 shows the CMD of NGC121 for stars on the PC of the WFPC2, obtained using DAOPHOT/ALLFRAME for the data reduction. The variable stars are marked by different symbols (see caption). Examples of light curves for RR Lyrae stars and DCs are shown in Figs. 2-3.\\\\ The conspicuous number of RR Lyrae stars detected in NGC121 confirms the ``true\" globular cluster nature of NGC121, a cluster that is at the young end of the Galactic globulars' age range. The apparent presence of both $\\delta$ Scuti and SX Phoenicis stars in NGC121 is an intriguing feature, that needs to be confirmed by a more detailed analysis of the light curves and a careful study of the contaminations by the SMC field stars." }, "0801/0801.3303_arXiv.txt": { "abstract": "A theoretical model framework of spherical symmetry is presented for a composite astrophysical system of two polytropic fluids coupled together by gravity to explore large-scale shocks and flow dynamics in clusters of galaxies or in globular clusters. The existence of such large-scale shocks in clusters of galaxies as inferred by high-resolution X-ray and radio imaging observations implies large-scale systematic flows that are beyond usual static models for clusters of galaxies. Here, we explore self-similar two-fluid flow solutions with shocks for a hot polytropic gas flow in a cluster of galaxies in the presence of a massive dark matter (DM) flow after the initiation of a gravitational core collapse or a central AGN activity or a large-scale merging process. In particular, the possibility of DM shocks or sharp jumps of mass density and of velocity dispersion in dark matter halo is discussed and such DM shocks might be detectable through gravitational lensing effects. To examine various plausible scenarios for clusters of galaxies, we describe three possible classes of shock flows within our model framework for different types of temperature, density and flow speed profiles. Depending upon sensible model parameters and shock locations, the hot ICM and DM halo may have various combinations of asymptotic behaviours of outflow, breeze, inflow, contraction or static envelopes at large radii at a given time. We refer to asymptotic outflows of hot ICM at large radii as the {\\it galaxy cluster wind}. As a result of such galaxy cluster winds and simultaneous contractions of DM halo during the course of galaxy cluster evolution, there would be less hot ICM within clusters of galaxies as compared to the average baryon fraction in the Universe. Physically, it is then expected that such `missing baryons' with lower temperatures reside in the periphery of galaxy clusters on much larger scales. Based on our model analysis, we also predict a limiting (the steepest) radial scaling form for mass density profiles of $r^{-3}$ within clusters of galaxies. ", "introduction": "Extensive X-ray observations have revealed that almost completely ionized hot gas medium permeates within clusters of galaxies with a typical temperature range of $\\sim 10^7-10^8$ K and a range of typical electron number density $\\sim 10^{-2}-10^{-4}\\hbox{ cm}^{-3}$ (e.g., Cavaliere \\& Fusco-Femiano 1978; Sarazin 1988; Fabian 1994). Clusters of galaxies are largely gravitationally bound systems on spatial scales of several Mpcs; together with the strong evidence of high velocity dispersions of galaxies (e.g., $\\sim 700-1000\\hbox{ km s}^{-1}$), hot X-ray emitting thermal electron gas, and gravitational lensing effects, we have realized the presence of massive dark matter halo within clusters of galaxies. Physical properties of galaxy clusters with static models of spherical symmetry have been extensively studied in the past (e.g., Lea 1975; Cavaliere \\& Fusco-Femiano 1976, 1978; Sarazin \\& Bahcall 1977; Sarazin 1988; Fabian 1994; Carilli \\& Taylor 2002; Voit 2005). In the past several years, high-resolution X-ray imaging observations have revealed density, temperature and pressure jumps in the hot intracluster medium (ICM) within clusters of galaxies (e.g., Fabian et al. 2003; Nulsen et al. 2005a; Nulsen et al. 2005b; McNamara et al. 2005), indicating that these structures are likely large-scale shocks rather than cold fronts (e.g., Sanders \\& Fabian 2006). Active galactic nuclei (AGNs) (e.g., Nulsen et al. 2005a; Nulsen et al. 2005b; McNamara et al. 2005) and merging galaxies (e.g., Markevitch \\& Vikhlinin 2001; Markevitch et al. 2002; Gabici \\& Blasi 2003; Markevitch et al. 2005) are proposed to be the driving force and energy source of these large-scale shocks. In addition, large-scale sound waves in clusters of galaxies have been proposed by Sanders \\& Fabian (2007) to explain the observed quasi-concentric ripples in surface brightness of X-ray emissions. Once shocks are identified in clusters of galaxies, there must be large-scale flows involved. In other words, these clusters of galaxies cannot be really static on large scales at least in the spatial region where shocks are presumably identified. Relative to the cluster centre, radial distances of shocks observed vary from tens of kpcs (e.g., McNamara et al. 2005; Nulsen et al. 2005b) to several Mpcs (e.g., A3667 in Rotteringer et al. 1997; A3376 in Bagchi et al. 2006; galaxy clusters A786, A2255, A2256 in Ensslin et al. 1998); at yet smaller radii around the centre, there could be emerging shocks that may not be easily identified. We take the point of view that these shocks are moving in hot ICM and we simply catch them at different epochs of evolution. These shocks may all be born somewhere around the central region and travel outwards to the locations we observe at the present epoch. For those clusters of galaxies without shock signatures, one possibility is that shocks have occurred in the distant past and have disappeared after their energies were dissipated during the process of propagation. So large-scale shocks and flows may well be common phenomena in clusters of galaxies. On much smaller scales compared to those of clusters of galaxies, X-rays have been also observed in globular clusters (e.g., Verbunt et al. 1984) and these emissions are interpreted by some as associated with flowing gas towards a black hole residing in the centre of globular clusters (e.g., Silk \\& Arons 1975). Moreover, outflows of gas materials from globular clusters have also been discussed in the literature (e.g., VandenBerg 1978). Fully relaxed globular clusters can be well treated as spherically symmetric (e.g., Harris \\& Racine 1979). One may view the collection of stars as one `fluid' and the tenuous gas as another fluid; these two fluids are coupled together by gravity on large scales. Here, we note globular clusters in passing and will mainly focus on large-scale self-similar dynamics for clusters of galaxies. With these two classes of astrophysical systems in mind, we develop a theoretical model framework of spherical symmetry to study dynamic behaviours of hot ICM and dark matter halo using the two-fluid approximation (e.g., Lou 2005), where the two polytropic fluids are coupled together by gravity. We should note that here the notion of a polytropic fluid is fairly general in the sense of specific entropy conservation along streamlines (Lou \\& Cao 2007). In the context of large-scale structure formation, extensive numerical works have been carried out to simulate the formation of galaxy clusters in the expanding universe, providing information for the hot ICM and dark matter halo (see, e.g., Bertschinger 1998 for a review on numerical simulations of structure formation in the universe). Evrard (1990) and Thomas \\& Couchman (1992) simulated properties (such as the number density and temperature profiles) of a hot gas in the presence of a dark matter halo. In Katz \\& White (1993) and Frenk et al. (1996), the radial cooling process is simulated. In particular, Evrard et al. (1994) simulated the formation of galaxies with two gravitationally coupled fluids representing dark matter halo and baryon matter, which is similar in essence to the approximation adopted in our semi-analytical model for clusters of galaxies but on much larger scales. Various self-similar solutions describing hydrodynamic processes of a single self-gravitational isothermal or polytropic gas under spherical symmetry have been investigated previously in contexts of star formation (e.g., Larson 1969a; Larson 1969b; Shu 1977; Hunter 1977; Shu et al. 1987; Lou \\& Shen 2004). Very recently, asymptotic behaviours of novel quasi-static solutions in a single polytropic gas sphere with self-gravity have been reported by Lou \\& Wang (2006, 2007) and was utilized to model rebound (MHD) shocks in supernovae (Wang \\& Lou 2007). For an astrophysical system of two fluids coupled by gravity, we can systematically extend these self-similar solutions, especially the new quasi-static solution, which may be used to describe behaviours of hot ICM and dark matter halo in clusters of galaxies. Except for the gravitational effect in the Newtonian sense, nothing else is known about dark matter particles at present. Using the coupled two-fluid model, we might be able to learn physical properties of dark matter halo through detectable diagnostics of hot ICM and of gravitational lensing effects. For clusters of galaxies, there is an outstanding problem of `missing baryons'. Extensive X-ray observations have indicated that the baryon mass fraction in clusters of galaxies is typically less than the prediction of primordial nucleosynthesis (e.g., Ettori \\& Fabian 1999; Ettori 2003; He et al. 2005; McGaugh 2007). This discrepancy becomes more difficult to reconcile in the cores of galaxy clusters (e.g., Sand et al. 2003). The best fit of cosmological parameters with tiny temperature fluctuations of the cosmic microwave background (CMB) radiation and large-scale structure clustering shows that relative to the critical mass density $\\rho_c$ in the universe, the mass density of baryon matter is $\\Omega_b=0.0224\\pm 0.0009 h^{-2}_{100}$ and the total matter density is $\\Omega_m=0.135^{+0.008}_{-0.009}h^{-2}_{100}$, where parameter $h_{100}$ is related to the Hubble constant $H_0$ by $H_0=100h_{100}$ km s$^{-1}$ Mpc$^{-1}$. Therefore, the mean cosmic baryon mass fraction is $f_b\\equiv\\Omega_b/\\Omega_m=0.166^{+0.012}_{-0.013}$ (e.g., He et al. 2005 and references therein). While there are different methods in determining the $f_b$ value, the cosmic baryon fraction $f_b$ is around 0.17 (e.g., McGaugh 2007). However in clusters of galaxies, the average gas (baryon) fraction inferred by two methods are about $0.107^{+0.028}_{-0.019}$ and $0.111^{+0.069}_{-0.063}$ (e.g., Ettori 2003). Others estimated that the baryon fraction $f_b$ observed in clusters of galaxies can be lower than the cosmic baryon fraction by about $10\\% - 20\\%$ at $z=0$ (e.g., He et al. 2005). Some even claimed that the value of $f_b$ can be lowered by as much as 30\\% (e.g., Ettori 2003). In conclusion, the baryon fraction in most clusters of galaxies are systematically lower than the average cosmic value $f_b$ except those highest estimates for gas (baryon) mass fraction in some clusters of galaxies (e.g., A426, A2142, RXJ1350; see Ettori 2003). To resolve this important issue, the notion of Warm-Hot Intergalactic Medium (WHIM) has been introduced (e.g., Cen \\& Ostriker 1999, 2006; Ettori 2003). In their opinion, the WHIM may actually exist within clusters of galaxies to account for the mass of `missing baryons', yet the WHIM cannot be detected at present because it does not emit X-rays. These results show that a significant fraction ($\\sim 40\\%-50\\%$) of the baryon component might be found in the form of WHIM in the temperature range of $T\\sim 10^{5-7}$ K (e.g., Cen \\& Ostriker 2006). As will be discussed in more details, this problem of `missing baryons' in our model scenario is a natural consequence of galaxy cluster winds, be it sustained or sporadic or be it stationary or dynamic during the evolution of galaxy clusters. These so-called `missing baryons' are blown away in the form of hot ICM and cool down gradually with time; with relatively low temperatures, they should mostly reside in the periphery of galaxy clusters and spread out in space on much larger scales. Meanwhile, the dark matter halo may contract within clusters of galaxies in our model. Therefore the mass fraction of baryons $f_b$ (i.e., the mass ratio of total baryons to the total gravitational mass inferred) would be lower than the initial value when a cluster of galaxies was born and started to evolve. The age of galaxy clusters is estimated to fall in the range of $\\sim 10^9\\ -\\ 10^{10}$ yr (e.g., Fabian 1994). As galaxy cluster winds may have existed since galaxy clusters were born, the timescale of galaxy cluster winds would be comparable to or somewhat less than this estimate. In Section 3, we show a few specific examples of numerical shock flow solutions in our model and estimate the loss of baryons within a timescale of $\\sim 10^9$ yr. As different behaviours of temperature profiles have been inferred from X-ray observations of galaxy clusters (Markevitch 1996 and Markevitch et al. 2005 for decreasing temperatures with increasing radius; Peres et al. 1998 and Sanders \\& Fabian 2006 for nearly constant temperatures in several galaxy clusters; McNamara et al. 2005 and Blanton et al. 2001 for increasing temperatures with increasing radius) and electron number densities are observed to fit a power law fairly well (e.g., Peres et al. 1998; Nulsen et al. 2005b), we shall take the specific entropy conservation along streamlines as the equation of state and see how well this may account for the various observed profiles of thermodynamic variables. By properly choosing model parameters in various regimes, we can describe properties of galaxy clusters to a considerable extent. This paper is structured as follows. The background and motivation of our model development is introduced in Section 1. Section 2 presents in order the basic formulation for the two-fluid model of spherical symmetry involving two polytropic fluids, self-similar transformation, asymptotic solutions at small and large $x$, singular surfaces and sonic critical curves, and shock conditions. In Section 3, we show numerical examples of quasi-static solutions for three different situations. The major results are summarized in Section 4. Finally we discuss our model results and numerical solutions in Section 5. Certain mathematical details are contained in Appendices A through G for the convenience of reference. ", "conclusions": "In this paper, we formulate a two-fluid model framework of spherical symmetry to explore dynamic behaviours of hot ICM and dark matter during the evolution of galaxy clusters. In this scenario, the hot ICM and dark matter halo are approximated as two polytropic `fluids' and are coupled by gravity. For both `fluids' in general situations, specific entropies are conserved along streamlines separately and are related to the enclosed masses. Quasi-static solutions for both `polytropic fluids' can be obtained and are adopted for sufficiently small radii around the central core region at a given time. In order to construct dimensionless quasi-static solutions in the regime of small $x$, we need to specify nine dimensionless parameters $\\{n,\\ \\gamma_1,\\ \\gamma_2,\\ \\kappa,\\ L_1,\\ L_2,\\ x_{ini},\\ x_{d,1},\\ x_{d,2}\\}$ where $x_{ini}$ needs to be properly chosen. Two additional dimensional parameters $K_2$ and $t$ need to be specified for dimensional solutions for a physical description. The ICM temperature is taken to be on the order of $\\sim 10^7-10^8$ K in typical clusters of galaxies and the electron number density is taken to be $\\sim 10^{-2}-10^{-4}$ cm$^{-3}$ in the typical radial range of kpcs to Mpcs. It is possible to construct different types of flow solutions with shocks in both fluids and with various asymptotic scaling features in flow speeds, mass densities, enclosed mass and temperatures. In particular, we can construct dynamic solutions for galaxy cluster winds and discuss the important problem of `missing baryons' during the evolution of galaxy clusters. There are several physical hypotheses in our two-fluid model to simplify the mathematical analysis. First, we take the dark matter halo as a kind of `fluid' for simplicity and explore this alternative theoretical possibility. Secondly, we assume the two-fluid system of galaxy clusters to be grossly spherically symmetric with a common centre and hope to catch major dynamic flow features on large scales. Thirdly, both `fluids' are assumed to be `polytropic' in the most general sense of entropy conservation along streamlines. In other words, the specific entropy distribution is not necessarily constant in space and time but are allowed to vary in $r$ and $t$ in general. Finally, dark matter interacts with the hot ICM only through gravity. Global semi-complete solutions can be constructed to pass through the two singular surfaces via shocks in both fluids. As large-scale ICM shocks have been identified observationally, there must be large-scale flows of hot ICM. As the dark matter halo and the hot ICM are coupled by gravity, there may be dark matter flows and the possibility of shocks in dark matter halo. Such dark matter shocks are characterized by drastic density jumps and sharp rises of velocity dispersions and may be detected by utilizing gravitational lensing effects. In our model framework, outflows of hot ICM in galaxy clusters actually form galaxy cluster winds, which is a systematic mechanism of reducing the baryon fraction $f_b$. In our model, the self-similar shocks travel outwards in both hot ICM and dark matter halo, respectively. The travel speeds of these shocks actually change with time for $n\\neq 1$ and their radial positions at present can vary from tens of kpcs to a few Mpc. Due to galaxy cluster winds, the baryon fraction $f_b$ in galaxy clusters can be $\\sim 15\\%-40\\%$ lower than the average cosmic baryon fraction in the Universe, which may account for the problem of `missing baryons' in clusters of galaxies. Physically, these `missing baryons' should reside in the periphery of galaxy clusters in the form of warm gas as results of unavoidable radiative cooling. Since the lower baryon fraction $f_b$ is a generic phenomenon in clusters of galaxies, we therefore suggest that galaxy cluster winds would be common and frequent during the evolution of galaxy clusters. The main features of our self-similar polytropic solution for two fluids coupled by gravity are summarized below. The radial profile of mass density at a given time is $\\rho_i\\propto r^{-2/n}$ in both fluids for either $r\\rightarrow 0^{+}$ or $r\\rightarrow+\\infty$ with $2/31$ and $x\\rightarrow+\\infty$, the radial flow velocities of both fluids become divergent for $H_1\\neq 0$ and $H_2\\neq 0$ in asymptotic solution (\\ref{largev1}) and (\\ref{largev2}). For $H_1=H_2=0$, the radial flow velocities of both fluids remain finite at large $x$ for $n>1$. The radial profile of temperature in the hot ICM is $T_2\\propto r^{2-2/n}$ for either $r\\rightarrow 0^{+}$ or $r\\rightarrow+\\infty$. By mass conservation, the enclosed mass is continuous across shocks." }, "0801/0801.4465_arXiv.txt": { "abstract": "Published data for large amplitude asymptotic giant branch variables in the Large Magellanic Cloud are re-analysed to establish the constants for an infrared ($K$) period-luminosity relation of the form: $M_K=\\rho[\\log P-2.38] + \\delta$. A slope of $\\rho=-3.51\\pm0.20$ and a zero point of $\\delta=-7.15\\pm0.06$ are found for oxygen-rich Miras (if a distance modulus of $18.39\\pm0.05$ is used for the LMC). Assuming this slope is applicable to Galactic Miras we discuss the zero-point for these stars using the revised {\\it Hipparcos} parallaxes together with published VLBI parallaxes for OH Masers and Miras in Globular Clusters. These result in a mean zero-point of $\\delta=-7.25\\pm0.07$ for O-rich Galactic Miras. The zero-point for Miras in the Galactic Bulge is not significantly different from this value. Carbon-rich stars are also discussed and provide results that are consistent with the above numbers, but with higher uncertainties. Within the uncertainties there is no evidence for a significant difference between the period-luminosity relation zero-points for systems with different metallicity. ", "introduction": "Large amplitude Asymptotic Giant Branch (AGB) variables (Miras) are important distance indicators for old and intermediate age populations. They are luminous, both bolometrically and in the near-infrared, and easily identified by their late spectral types (Me, Ce, Se or very rarely Ke), large amplitudes ($\\Delta V > 2.5$ mag, $\\Delta K > 0.4$ mag) and long periods ($100 \\lsim P \\lsim 1000$). The increasing use of adaptive optics on large telescopes at near-infrared wavelengths to study stellar populations (e.g. Da Costa 2004) at large distances will require confidence in the calibration of the AGB variables as distance indicators. Alternatively, if the distance is known the luminosities of the brightest AGB stars provide insight into the intermediate age populations, e.g. Menzies et al. (2008). Wood et al. (1999) demonstrated that, within the LMC, AGB variables fall on a series of, approximately parallel, period-luminosity (PL) relations at $K$, (see also Cioni et al. 2001, Ita et al. 2004, Fraser et al. 2005, Soszy\\`nski et al. 2007). The large amplitude variables, i.e. the Miras, however, lie only on a singe PL($K$) relation. The existence of a Mira PL($K$) relation has been known for some while (Feast et al. 1989; Hughes \\& Wood 1990) and is now generally thought to represent the relation for fundamental pulsation. While some of the low amplitude variables also pulsate in the fundamental, others show various overtones. Low amplitude variables are of limited use for distance scale studies as it is not simple to establish upon which of the various PL relations they lie. Complications do arise at periods in excess of about 400 d where some Miras in the LMC have higher luminosities, possibly as a consequence of hot bottom burning (e.g. Whitelock et al. 2003). Feast (2004) provides a recent discussion of AGB stars as distance indicators and of the zero-point of the PL relation. Following a brief re-examination of the Mira PL($K$) relation in the LMC, we take advantage of the new analysis of the {\\it Hipparcos} data (van Leeuwen 2007a and 2007b, see also van Leeuwen 2005 and van Leeuwen \\& Fantino 2005) to re-examine the distance scale for large amplitude AGB variables within the Galaxy. Analysis of the original {\\it Hipparcos} data was discussed by van Leeuwen et al. (1997), Whitelock \\& Feast (2000 - henceforth Paper~I) and Knapp et al. (2003). In the last part of this paper we put together all the available information on Mira parallaxes to establish the best value for the zero-point of the Mira PL($K$) relation for the Galaxy and compare this with values for elsewhere. \\begin{table} \\centering \\caption{LMC variables used to establish the slope for the Mira PL($K$) relation}\\label{lmcdata} \\begin{center} \\begin{tabular}{lccc} \\hline \\multicolumn{1}{c}{name}& P & $K$ & note\\\\ & \\multicolumn{1}{c}{(d)} & \\multicolumn{1}{c}{(mag)} \\\\ \\hline \\multicolumn{4}{c}{O-rich stars}\\\\ 0517--6551 & 116 & 12.25 & \\\\ C38 & 130 & 12.12 & \\\\ 0512--6559 & 141 & 12.13 & \\\\ W132 & 156 & 11.67 & \\\\ 0526--6754 & 157 & 11.79 & \\\\ W151 & 174 & 11.74 & \\\\ W148 & 183 & 11.82 & \\\\ W158 & 194 & 11.77 & \\\\ 0528--6531 & 195 & 11.48 & \\\\ C11 & 202 & 11.51 & \\\\ GR13 & 202 & 11.59 & \\\\ SHV05220--7012 & 205 & 11.73 & 1\\\\ 0507--6639 & 208 & 11.57 & \\\\ C20 & 210 & 11.54 & \\\\ W77 & 213 & 11.25 & S\\\\ R120 & 217 & 11.38 & \\\\ W94 & 220 & 11.28 & \\\\ W74 & 231 & 11.49 & \\\\ WBP74 & 233 & 11.50 & 1\\\\ W1 & 235 & 11.48 & \\\\ W140 & 243 & 11.19 & \\\\ 0533--6807 & 247 & 11.38 & \\\\ R141 & 258 & 10.99 & \\\\ R110 & 261 & 11.29 & \\\\ W48 & 279 & 10.99 & \\\\ 0537--6607 & 284 & 11.02 & \\\\ 0505--6657 & 307 & 10.67 & \\\\ 0524--6543 & 315 & 10.71 & \\\\ W126 & 318 & 10.89 & K\\\\ SHV05305--7022 & 362 & 10.57 & \\\\ R105 & 413 & 10.33 & \\\\ \\multicolumn{4}{c}{C-rich stars}\\\\ 0530--6437 & 157 & 12.08 & \\\\ 0515--6617 & 226 & 11.16 & \\\\ 0528--6520 & 229 & 11.08 & \\\\ 0520--6528 & 233 & 11.28 & \\\\ 0519--6454 & 242 & 11.09 & \\\\ W220 & 281 & 10.83 & \\\\ 0529--6759 & 283 & 10.91 & \\\\ 0515--6451 & 284 & 10.81 & \\\\ SHV05027--6924 & 298 & 10.82 & \\\\ 0514--6605 & 308 & 10.64 & \\\\ 0534--6531 & 308 & 10.98 & \\\\ 0529--6739 & 319 & 10.60 & \\\\ 0502--6711 & 322 & 10.53 & \\\\ C7 & 327 & 10.69 & \\\\ 0541--6631 & 342 & 10.50 & \\\\ R153 & 347 & 10.52 & \\\\ WBP14 & 351 & 10.62 & \\\\ W103 & 363 & 10.78 & \\\\ 0515--6438 & 365 & 10.90 & \\\\ 0537--6740 & 367 & 10.47 & \\\\ SHV05003--6817 & 369 & 10.58 & \\\\ SHV05260--7011 & 373 & 10.54 & \\\\ \\hline \\multicolumn{4}{l}{notes 1. Period taken from MACHO;}\\\\ \\multicolumn{4}{l}{S and K are spectral types}\\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\begin{figure*} \\includegraphics[width=16cm]{lmcpl.ps} \\caption{The PL($K$) relation for Mira variables in the LMC. Solid and open symbols are O- and C-rich respectively. The dashed line is the fit to the O-rich stars from Table~\\ref{lmcpl} assuming a distance modulus for the LMC of 18.39 mag, while the dotted line is the relation for C-stars.} \\label{fig_lmcpl} \\end{figure*} ", "conclusions": "In Paper~I (section 2.2) we noted a caveat with regard to temporal changes in the light distribution across the stellar disk as problematic for the interpretation of the parallaxes. This is particularly so because the angular diameters of the Miras are two or three times larger than their parallaxes. More recent work (e.g. Ragland et al. 2006; Woodruff et al. 2008) confirms that the diameters are large, variable and non-uniform. These same references further highlight the disagreement between observation and theory, emphasizing our very limited understanding of the atmospheres of these stars or their variations. This is clearly a real problem, but the agreement on the distance scale achieved by the different methods summarized above may indicate that the net effect is small. As noted, the various results discussed above are in good agreement with each other. UX Cyg, is brighter than we would expect from the PL($K$) relation and this may be because of hot bottom burning. We take a simple mean of solutions 10, 25, 26 from Table~\\ref{zp}, to set a mean value of $\\delta = -7.25 \\pm 0.07$ for the O-rich Mira variables in the Galaxy. This is in reasonable agreement with the LMC value of $\\delta = -7.15 \\pm 0.06$. Solution 24, for the C-rich Miras, of $\\delta = -7.18 \\pm 0.37$ is in agreement with the above result for O-rich Miras and with the LMC value for C-stars of $\\delta = -7.24 \\pm 0.07$ This supports an identical PL($K$) relation for O- and C-rich variables. The O-rich result represents the best value for large amplitude variables, but it should be used with caution on stars with periods in excess of 400 days. The $\\rm H_2O$ parallaxes offer the most likely significant improvement in this result in the near future." }, "0801/0801.1847_arXiv.txt": { "abstract": "Standard calculations suggest that the entropy of our universe is dominated by black holes, whose entropy is of order their area in Planck units, although they comprise only a tiny fraction of its total energy. Statistical entropy is the logarithm of the number of microstates consistent with the observed macroscopic properties of a system, hence a measure of uncertainty about its precise state. Therefore, assuming unitarity in black hole evaporation, the standard results suggest that the largest uncertainty in the future quantum state of the universe is due to the Hawking radiation from evaporating black holes. However, the entropy of the matter precursors to astrophysical black holes is enormously less than that given by area entropy. If unitarity relates the future radiation states to the black hole precursor states, then the standard results are highly misleading, at least for an observer that can differentiate the individual states of the Hawking radiation. ", "introduction": " ", "conclusions": "" }, "0801/0801.1076_arXiv.txt": { "abstract": "We report on next phase of our study of rotating accretion flows onto black holes. We consider hydrodynamical (HD) accretion flows with a spherically symmetric density distribution at the outer boundary but with spherical symmetry broken by the introduction of a small, latitude-dependent angular momentum. We study accretion flows by means of numerical two-dimensional, axisymmetric, HD simulations for variety of the adiabatic index, $\\gamma$ and the gas temperature at infinity, $c_\\infty$. Our work is an extension of work done by Proga \\& Begelman who consider models for only $\\gamma=5/3$. Our main result is that the flow properties such as the topology of the sonic surface and time behavior strongly depend on $\\gamma$ but little on $c_\\infty$. In particular, for $1 < \\gamma < 5/3$, the mass accretion rate shows large amplitude, slow time-variability which is a result of mixing between slow and fast rotating gas. This temporal behavior differs significantly from that in models with $\\gamma\\simless 5/3$ where the accretion rate is relatively constant and from that in models with $\\gamma\\simgreat 1$ where the accretion exhibits small amplitude quasi-periodic oscillations. The key parameter responsible for the differences is the sound speed of the accretion flow which in turn determines whether the flow is dominated by gas pressure, radiation pressure or rotation. Despite these differences the time-averaged mass accretion rate in units of the corresponding Bondi rate is a weak function of $\\gamma$ and $c_\\infty$. ", "introduction": "Many types of astrophysical objects are powered by black hole (BH) accretion. Radiation energy produced by accretion can be very high and can explain such dramatic phenomena as quasars, powerful radio galaxies, X-ray binaries, and gamma ray bursts (GRBs). However, BH accretion does not always result in high radiative output. This is true for both stellar BH and super massive black holes (SMBH). In particular, objects with SMBH appear to spend statistically most of their time in an inactive phase. Inactive SMBHs are not something that one would expect, because these black holes are embedded in the relatively dense environments of galactic nuclei. Therefore it is natural to suppose that the gravity due to an SMBH will draw in matter at high rates, leading to a high system luminosity. Monitoring of X-ray binaries with stellar BHs reveals that these objects often exhibit large time variability in the total energy output in the spectral energy distribution. Then one of the main goals of any theory of BH accretion is two explain why accretion proceeds through very different modes. Generally, the radiative output from accretion depends on the mass accretion rate $\\MDOT_a$ and an efficiency factor, $\\eta$. In previous papers of this series (Proga \\& Begelman 2003a,b, hereafter PB03a and PB03b, respectively), we studied how physical conditions at large distances from SMBH affect $\\MDOT_a$ in the so-called radiatively inefficient accretion flows (RIAF). RIAF with very low $\\eta$ and also low $\\MDOT_a$ have been proposed to explain very low radiative luminosities in systems as such Sgr~A* (e.g., Ichimaru 1977; Rees et al. 1982; Narayan \\& Yi 1994, 1995; Abramowicz et al. 1995; Blandford \\& Begelman 1999; Sharma et al. 2007 and reference therein). In PB03a, we addressed the issue of how $\\MDOT_a$ depends on the distribution of specific angular momentum, $l$ at large radii assuming that the adiabatic index, $\\gamma=5/3$. For high $l$, corresponding to the circularization radius larger than the last stable orbit, gas cannot directly accrete onto a BH, unless some physical mechanism like e.g., viscosity or magnetic fields, or both, transports angular momentum towards. For inviscid accretion flow the matter with too high $l$, either flows outward due to gas pressure and centrifugal forces or accumulates close the central accretor (e.g. Hawley et al. 1984a,b,; Clarke et al. 1985; Chen et al 1997; PB03a; Janiuk $\\&$ Proga 2007). The simulations presented by PB03a illustrate a general flow pattern with an inflow in the polar funnel, and an equatorial outflow (Fig.~1 there). Such flow pattern is induced by the angular momentum distribution simply because rotating gas tend to converge toward the equator. If $l$ is high, the gas forms a subsonic very dynamic torus with gas flowing out. The mass accretion rate can be significantly smaller compared to corresponding the Bondi rate. In this work, we address the problem how the properties of the accretion flow presented in PB03a, change when different micro-physical properties of the flow are assumed. In particular, we explore effects of changing $\\gamma$ in the polytropic equation of state. In reality, accretion flows with different $\\gamma$'s, may correspond to different types of objects or different phases of activity. For example, in the weakly active galaxies like e.g. Sgr A*, $\\gamma$ is usually considered to be around 5/3, because these flows are believed to be radiatively inefficient and gas pressure dominated. On the other hand, the large energetic output observed in GRBs indicates that an accretion flow must be radiation pressure dominated (e.g., Meszaros 2006) and a relativistic equation of state is required with $\\gamma$=4/3. Another example are protogalactic disks, which are sometimes considered as being formed by isothermal accretion flow ($\\gamma\\approx 1$; e.g., Mo et al. 1998). Also, the so-called `high' and `low' accretion states in the X-ray binary systems, may reflect physical proprieties changing in time (for a review see Done et al. 2007). Our work is a straightforward extension of PB03a's work, who showed results for $\\gamma$=5/3. We extend their models by considering the flows, for $\\gamma$ ranging between 1 and 5/3. We also perform additional calculations with sound speed at infinity $c_{s,\\infty}$ much smaller than that considered by PB03a allowing us to model flows with the Bondi radius as large as that estimated in real systems, for example in Sgr~A*. Taking advantage of faster computers, we perform simulations with higher resolution in the $\\theta$ direction, and with much larger computational domain in comparison to PB03a. We keep other model parameters such as angular momentum distribution as in PB03a. For $\\gamma=5/3$, our new simulations are consistent with those presented by PB03a. However, we find that $\\MDOT_a$, the flow structure, defined by the gas density and angular momentum distribution, and the sonic surface topology depend on $\\gamma$. In particular,for $\\gamma=5/3$, $\\MDOT_a$ is nearly constant whereas for $\\gamma=4/3$, $\\MDOT_a$ shows stochastic, large-amplitude time-variability. For $\\gamma=1.01$, $\\MDOT_a$ shows small-amplitude periodic changes. In Sec.2, we describe a general set-up of our simulations. In Sec.3, we present our results. In the last section, we discuss the results, and relate them to results found in previous studies. ", "conclusions": "We report on next phase of our study of rotating accretion flows onto black holes. We consider hydrodynamical (HD) accretion flows with a spherically symmetric density distribution at the outer boundary but with spherical symmetry broken by the introduction of a small, latitude-dependent angular momentum. We study accretion flows by means of numerical two-dimensional, axisymmetric, HD simulations for variety of the adiabatic index, $\\gamma$ and the gas temperature at infinity, $c_\\infty$. Our work is an extension of work done by PB03a who consider models for only $\\gamma=5/3$. We also reran some models from PB03a with higher resolution. The latter are fully consistent with lower resolution PB03a's runs. Our main result is that the flow properties such as the topology of the sonic surface and time behavior strongly depend on $\\gamma$ but little on $c_\\infty$. In particular, for $1 < \\gamma < 5/3$, the mass accretion rate shows large amplitude, slow time-variability which is a result of mixing between slow and fast rotating gas. This temporal behavior differs significantly from that in models with $\\gamma\\approx 5/3$ where the accretion rate is relatively constant and from that in models with $\\gamma\\approx 1$ where the accretion exhibits small amplitude quasi-periodic oscillations. The key parameter responsible for the differences is the sound speed of the accretion flow which in turn determines the strength of shocks and whether the flow is dominated but gas pressure ($\\gamma\\simless 5/3$), radiation pressure ($1 < \\gamma < 5/3$) , or rotation ($\\gamma\\simgreat 1$). Despite these differences, the time-averaged mass accretion rate in units of the corresponding Bondi rate is a weak function of $\\gamma$ and $c_\\infty$. We realize that our simulations do not capture several physical processes that may be important in real systems (e.g., magnetic fields, radiative cooling and heating, energy dissipation). However, our simulations are done within a general framework so that some of these processes can be straightforwardly added (e.g., magnetic fields as in PB03b, or radiative processes as in Proga 2007). Therefore, this work could serve as a good reference point to analyze and interpret more complete and complex simulations (e.g., we already reran some of the models from this paper including magnetic fields, Moscibrodzka and Proga, in preparation). ACKNOWLEDGMENTS: We thank Bozena Czerny and Marek Abramowicz for very useful comments. We acknowledge support provided by the Chandra award TM7-8008X issued by the Chandra X-Ray Observatory Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of NASA under contract NAS8-39073. M.M also acknowledges supported in part by grant 1P03D~008~29 of the Polish State Committee for Scientific Research (KBN). While D.P. acknowledges support from NASA under ATP grant NNG05GB68G. \\newpage" }, "0801/0801.0593_arXiv.txt": { "abstract": "Simple one--zone homogeneous synchrotron self--Compton models have severe difficulties is explaining the TeV emission observed in the radiogalaxy M87. Also the site the TeV emission region is uncertain: it could be the unresolved jet close to the nucleus, analogously to what proposed for blazars, or an active knot, called HST--1, tens of parsec away. We explore the possibility that the TeV emission of M87 is produced in the misaligned subpc scale jet. We base our modelling on a structured jet, with a fast spine surrounded by a slower layer. In this context the main site responsible for the emission of the TeV radiation is the layer, while the (debeamed) spine accounts for the emission from the radio to the GeV band: therefore we expect a more complex correlation with the TeV component than that expected in one--zone scenarios, in which both components are produced by the same region. Observed from small angles, the spine would dominate the emission, with an overall Spectral Energy Distribution close to those of BL Lac objects with a synchrotron peak located at low energy (LBLs). ", "introduction": "An increasing number of extragalactic objects is detected at energies greater than 100 GeV. The list of published sources comprises 20 objects\\footnote{see {\\tt http://www.mppmu.mpg.de/$\\sim$rwagner/sources}}: as expected (e.g., Costamante \\& Ghisellini 2002), the bulk of them (17) belongs to the class of Highly Peaked BL Lac objects. In fact, the position of the synchrotron peak, usually located in (or close to) the X--ray band assures the existence, in these sources, of electrons with extremely large Lorentz factors ($\\gamma=10^5$--$10^6$), a key ingredient for the emission in the TeV band via inverse Compton (IC) scattering. The remaining three sources are BL Lac itself, which belongs to the LBL class, (Albert et al. 2007), 3C279, a FSRQs, (Teshima et al. 2007) and the nearby (16 Mpc, Tonry 1991) radiogalaxy M87 (Aharonian et al. 2003, 2006). Already suggested as a possible source of high-energy radiation (e.g. Bai \\& Lee 2001) based on its similarity with BL Lac objects (Tsvetanov et al. 1998), M87 was discovered as a TeV source by the HEGRA array of Cherenkov telescopes (Aharonian et al. 2003) and it has been recently confirmed by H.E.S.S. (Aharonian et al. 2006). Due to the limited spatial resolution it is not possible to identify the emission region of this radiation. However, the sensitivity of H.E.S.S. allowed the detection of variability on short timescales ($\\sim$ days), suggesting a compact emission region. Three main sites have been proposed as emission regions for the TeV radiation: the resolved jet (in particular the so--called knot HST--1); the unresolved base of the jet (in analogy with blazars); the vicinity of the supermassive ($M=3\\times 10^9$ $M_{\\odot}$, Marconi et al. 1997) black hole (BH). HST--1 is a quite interesting jet feature located at 60 pc (projected) from the core of of M87. Superluminal motions with apparent speeds up to $\\sim 6 c$ were observed with {\\it HST} (Biretta et al. 1999). Continuous multifrequency coverage has shown spectacular activity from this knot in the past years (Perlman et al. 2003; Harris et al. 2006, 2007; Cheung et al. 2007). In particular, the X--ray flux increases by a factor of $\\sim 30$ from 2000 to 2005 and since then it is steadily decreasing. These variations are accompanied by similar changes in the optical and in the radio bands and by the appearance of new superluminal components. This extreme phenomenology is described by Stawarz et al. (2006) on the assumption that HST-1 marks the recollimation shock of the jet. As such, HST--1 is thought to be a rather efficient particle accelerator and thus a possible source of intense TeV radiation. However, the small variability timescales observed by H.E.S.S. put strong limits to the size of the source ($R < c \\Delta t \\, \\delta \\simeq 5\\times 10^{15} \\delta$ cm, where $\\delta=[\\Gamma(1-\\beta \\cos\\theta)]^{-1}$ is the relativistic Doppler factor, $\\theta$ is the viewing angle and $\\Delta t\\sim 2$ days), which seems difficult to accomplish in this scenario (e.g. Neronov \\& Aharonian 2007; Levinson 2007). The rapid variability could be easily reproduced if the TeV emission is produced close to the base of the jet, in the region associated to the emission from blazars. However, standard one--zone leptonic models face difficulties in reproducing the observed spectral energy distribution (SED) of the core including the TeV data. A possible solution advocates that the emission comes from a decelerating jet (Georganopoulos et al. 2005) . Alternatively, the emission can be produced by high--energy protons (Reimer et al. 2004). Going further in towards the BH, TeV photons could be emitted through IC by relativistic pairs produced by electromagnetic cascades in the BH magnetosphere (Neronov \\& Aharonian 2007). The determination of the site producing the high-energy emission is rather important. Indeed, if the emission site will be eventually identified as knot HST-1, this would have a broad impact on the current view of relativistic jets (see the discussion in Cheung et al. 2007). It is thus extremely important to investigate whether the standard view (or minimal variations of it) for the inner jet is able reproduce the observed phenomenology. If not, it would be mandatory to explore the other alternatives. As discussed in Ghisellini, Tavecchio \\& Chiaberge (2005, hereafter Paper I), there is compelling evidence supporting the view that jets of BL Lac objects are structured at blazar scales (at distances $\\sim 10^{17}$ cm from the BH), with a fast core ({\\it spine}) surrounded by a slower sheet ({\\it layer}). In Paper I we showed that the radiative coupling between the spine and the layer can naturally account for the deceleration inferred for the jet between the blazar and the VLBI scale. Furthermore, one component sees the radiation of the other boosted, since the relative velocity can be relativistic. This enhances the IC emission of both the spine and the layer. A direct consequence of such a structure is that while at small viewing angle (as in the case of blazars) the emission is dominated by the boosted spine emission, at large observing angles ($\\theta>45^\\circ$, typical for radio--galaxies) the emission from the spine would be suppressed, while the layer, characterised by a broader beaming cone, could substantially contribute to (and sometimes dominate) the overall emission. At intermediate angles both components can significantly contribute. Following this line, in this letter we explore an alternative interpretation for the TeV emission of M87, assuming that the TeV radiation is produced in the layer of the misaligned jet. We use $H_{\\rm 0}\\rm =70\\; km\\; s^{-1}\\; Mpc^{-1} $, $\\Omega_{\\Lambda}=0.7$, $\\Omega_{\\rm M} = 0.3$. ", "conclusions": "\\label{sec4} We propose that the TeV emission detected from M87 originates in the structured jet at blazar scale. The most likely possibility is that the emission from the spine reproduces the low energy component, from radio to X--rays, while the layer contributes to the bulk of the TeV radiation. With this choice, the beamed counterpart of M87 observed at small angle would have a SED closely resembling those of Low--Peaked BL Lac objects. Correspondingly the spine is characterised by physical parameters close to those usually inferred for those sources ($\\gamma _b\\sim 10^3$, $B\\sim 1$ G). The layer, instead, would be characterised by a low magnetic field ($B\\sim 0.1$ G) and large peak Lorentz factors ($\\gamma _b\\sim 10^6$). It is possible that particles in the layer are energized by turbulent acceleration (e.g. Stawarz \\& Ostrowski 2002). In this case the expected distribution would have a pile-up at the energy where acceleration and loss processes are in equilibrium (e.g. Katarzynski et al. 2006). Suggestively, in our conditions (losses dominated by IC) the expected equilibrium Lorentz factor would be close to $\\gamma \\sim 10^6$ for $B\\sim 0.1$ G. Note that our conclusions are rather different from that of Georganopoulos et al. (2005) who discussed a similar model relying on the possibility that the jet experiences a strong deceleration at the blazar scale. In their model the high-energy emission is produced in the innermost and faster portion of the jet, while the slow external portions mainly contributes at low frequency. Under these conditions the beamed counterpart of M87 would be a TeV BL Lac. However, due to the different beaming patterns or the synchrotron and IC components (analogously to our case), the resulting spectrum would be characterised by a strong dominance of the TeV component with respect to the X--ray one, contrary to what is usually observed in the known TeV BL Lacs. In our model the optical and the X--rays are produced by a region different from that responsible for the TeV emission. Therefore a strict correlation between low energy and TeV emission is not directly required (even if it is possible). Analogously, the MeV-GeV emission should originate in the spine. Therefore we expect that the emission possibly detected by {\\it GLAST} will not exactly follow the TeV component. A potential weak point of our model concerns the slope of the TeV spectrum. The slope is mainly dictated by the absorption of TeV photons by the dense optical radiation field and, in this sense, it is rather robust and does not strongly depend on the underlying slope of the intrinsic TeV spectrum. In particular, hard spectra such as that recorded in 2005 are quite difficult to achieve in this scheme. A possibility to avoid important absorption of gamma-rays would be to enlarge the emission region, hence diluting the target photon density and decreasing the absorption optical depth. However the increase of the source size is limited by the observed short variability timescales. Further observations with increased sensitivity could provide stringent constraints to our interpretation." }, "0801/0801.2769_arXiv.txt": { "abstract": "We measure the dependence of the AGN fraction on local environment at $z\\sim1$, using spectroscopic data taken from the DEEP2 Galaxy Redshift Survey, and Chandra X-ray data from the All-Wavelength Extended Groth Strip International Survey (AEGIS). To provide a clean sample of AGN we restrict our analysis to the red sequence population; this also reduces additional colour--environment correlations. We find evidence that high redshift LINERs in DEEP2 tend to favour higher density environments relative to the red population from which they are drawn. In contrast, Seyferts and X-ray selected AGN at $z\\sim1$ show little (or no) environmental dependencies within the same underlying population. We compare these results with a sample of local AGN drawn from the SDSS. Contrary to the high redshift behaviour, we find that both LINERs and Seyferts in the SDSS show a slowly declining red sequence AGN fraction towards high density environments. Interestingly, at $z\\sim1$ red sequence Seyferts and LINERs are approximately equally abundant. By $z\\sim0$, however, the red Seyfert population has declined relative to the LINER population by over a factor of $7$. We speculate on possible interpretations of our results. ", "introduction": "\\label{sec:intro} Both active galactic nuclei (AGN) and local environment play key roles in shaping galaxy evolution. It is now understood that AGN are those nuclei in galaxies that emit radiation powered by accretion onto a supermassive black hole. Although this realisation has proved useful for explaining many observed characteristics of these active objects, there are still many unsolved problems, especially related to the physics of the accretion process itself. In the recent years much effort has been invested in studying the global properties of AGN as a unique population in the context of galaxy formation. In this work, we focus on a fundamental question: the dependence of the fraction of galaxies that have AGN on the density of the local environment at $z\\sim1$, and the evolution of this dependence to $z\\sim0$. At low redshift, many authors have investigated various correlations between \\textit{galaxy properties} and environment. It is now well established that there exists a relationship between morphology and density (\\citealt{Oemler1974} and \\citealt{Dressler1980}), in that star-forming disk-dominated galaxies tend to inhabit less dense regions of the Universe than ``quiescent'' or inactive elliptical galaxies. Moreover, additional (and related) dependencies with environment have been found, such as with stellar mass, luminosity, colour, recent and past star formation, star formation quenching, surface brightness, and concentration (to name but a few) \\citep[e.g.][]{Kauffmann2004, Balogh2004, Hogg2004, Blanton2005, Bundy2006}. In this scenario of entangled correlations it is useful to investigate the dependence of AGN properties on the local environment, especially since AGN are believed to play an important part in shaping galaxy evolution. This has sometimes been a rather controversial issue. In the local Universe, \\cite{Miller2003} found no dependence on environment of the fraction of spectroscopically selected AGN, using the SDSS early data release. This result is in good agreement with \\cite{Sorrentino2006} who used the much larger SDSS DR4. However, many other authors claim the existence of a strong link between nuclear activity and environment, at least for specific AGN types. \\cite{Kauffmann2004} found that intermediate luminosity optically selected AGN (Seyfert IIs) favoured underdense environments, while low-luminosity optically selected AGN (Low-Ionization Nuclear Emission-line Regions; hereafter, LINERs) showed no density dependence, within the SDSS DR1. Similarly, lower-luminosity AGN were found to have a higher clustering amplitude than high-luminosity AGN by \\cite{Wake2004} and \\cite{Constantin2006a}. Radio-loud AGN have been noted to reside preferentially in mid-to-high density regions and tend to avoid underdense environments \\citep{Zirbel1997, Best2004}. At high redshift the study of both galaxies and AGN, and their relation to the environment, has been restricted by the lack of adequate data. Only in recent years, with the emergence of quality large-scale probes of the high redshift galaxy population, such as the DEEP2 Galaxy Redshift Survey \\citep{Davis2003} or the VIMOS-VLT Deep Survey \\cite[VVDS,][]{LeFevre2003}, have we reached the stage where we can begin to measure the statistics of galaxy evolution in some detail. Using DEEP2, \\cite{Cooper2006} found that the many of the low redshift galaxy correlations with environment are already in place at $z\\!\\sim\\!1$. However important differences exist. The colour-density relation, for instance, tends to weaken towards higher redshifts \\citep{Cooper2007a,Cucciati2006}. Also, bright blue galaxies are found, on average, in much denser regions than at low redshift. Such a population inverts the local star formation-density relation in overdense environments \\citep{Cooper2007b,Elbaz2007}. This inversion may be an early phase in a galaxy's transition onto the red sequence through the process of star formation quenching. The truncation of star formation in massive galaxies is believed to be tightly connected with nuclear activity \\citep[see e.g.][for more information]{Croton2006, Bower2006}. Further investigation reveals that post-starburst (aka. K+A or E+A) galaxies \\citep[e.g.][]{DresslerGunn1983} are galaxies ``caught in the act'' of quenching and are in transit to the red sequence. These predominantly ``green valley'' objects reside in similar environments to regular star forming galaxies (\\citealt{Hogg2006, Nolan2007}; Yan et al. 2007 in prep.) supporting the picture that star formation precedes AGN-triggered quenching, which precedes retirement onto the red sequence. \\cite{Georgakakis2007} were one of the first to study the environments of X-ray selected AGN at $z\\sim1$ using a sample of 58 sources drawn from the All-Wavelength Extended Groth Strip International Survey (AEGIS, \\citealt{Davis2007}). The authors found that these galaxies avoided underdense regions with a high level of confidence. \\cite{Nandra2007} show that the same AGN reside in host galaxies that populate from the top of the blue cloud to the red sequence in colour-magnitude space. They speculate that such AGN may be the mechanism through which a galaxy stays red. Similar ideas have become a popular feature of many galaxy formation models that implement lower luminosity (i.e. non-quasar) AGN to suppress the supply of cooling gas to a galaxy, hence quenching star formation through a process of ``starvation'' \\citep[e.g.][]{Croton2006, Bower2006}. In this work we study the environmental dependence of nuclear activity in red sequence galaxies within a carefully chosen sample of both X-ray and optically selected AGN, drawn from the AEGIS Chandra catalogue and the DEEP2 Galaxy Redshift Survey, respectively. Our paper is organised as follows. In Section~\\ref{sec:survey} we describe our AGN selection. In Section~\\ref{sec:results} we present our main result: the AGN fraction of red sequence galaxies at $z\\sim1$ as a function of environment for three types of AGN (LINERs, Seyferts and X-ray selected). We undertake a comparison between our high-z results and those derived from a low-z sample drawn from the SDSS in Section~\\ref{sec:sdss}. Finally, in Sections~\\ref{sec:discussion} and \\ref{sec:summary} we provide a discussion and brief summary. Throughout, unless otherwise stated, we assume a standard $\\Lambda$CDM concordance cosmology, with $\\Omega_m=0.3$, $\\Omega_\\Lambda=0.7$, $w=-1$, and $h=1$. In addition, we use AB magnitudes unless otherwise stated. \\begin{figure*} \\begin{center} \\begin{tabular}{c} \\includegraphics[scale=0.7]{./figures/FIG_paper1_2.ps} \\end{tabular} \\end{center} \\caption{Two panels that show our AGN selection of Seyferts and LINERS within the DEEP2. The left panel plots [OII] EW versus ${\\rm H{\\beta}}$ EW for objects with accurate redshifts ($Q\\ge3$), $\\delta_{3}$ environment measures, and covered [OII], [OIII] and ${\\rm H{\\beta}}$ (grey points). LINERs (black points) are selected using the empirical demarcation of Equation~\\ref{eqn:liners_selection} along with the colour cut defined by Equation~\\ref{eqn:color_deep2}. The right panel shows the line ratio ${\\rm [OIII]/ H{\\beta}}$ plotted against $(U-B)$ rest-frame colour for the same DEEP2 sample (grey points). Seyferts (black points) are selected to have ${\\rm [OIII]/H{\\beta}\\ge 3}$ and rest-frame colour $(U\\!-\\!B)>0.8$, as denoted by the horizontal and vertical lines. See Section~\\ref{sec:optical} for further details.} \\label{fig:OSS_selection} \\end{figure*} ", "conclusions": "\\label{sec:discussion} \\subsection{Previous measures of AGN and environment} It is difficult to make direct comparisons of our results with previously published works. This is because past environment studies have tended to focus on the AGN fraction of \\emph{all} colours of host galaxies, and also to mix both LINER and Seyfert classes into a combined AGN population. Our selection is restricted to the red sequence only (and also green valley), which allows us to compare high and low redshift populations and also study the AGN--environment connection without the colour--environment correlation. Locally, a number of SDSS measures of AGN and environment have been made. Using the SDSS early data release, \\cite{Miller2003} found no dependence on environment for the spectroscopically selected AGN fraction in a sample of 4921 objects. Specifically, the authors report no statistically significant decrease in the AGN fraction in the densest regions, although their densest points visually suggest such a trend. This result is broadly consistent with the results of both LINERs and Seyferts in Figure~\\ref{fig:SDSS_result}, even though we only consider red sequence objects. \\cite{Kauffmann2003} also found little environment dependence of the overall fraction of detected AGN in a sample drawn from the SDSS DR1. However, they do report different behaviour when the sample is broken into strong AGN ($\\log L[{\\rm OIII}]>7$, ``Seyferts'') and weak AGN ($\\log L[{\\rm OIII}]<7$, ``LINERs''). For Seyfert they find a significant preference for low-density environments, especially when hosted by more massive galaxies. This is consistent with our SDSS findings in Figure~\\ref{fig:SDSS_result} and different to what we find at $z\\sim1$ in the DEEP2 fields. For LINERs, \\citeauthor{Kauffmann2003} measure little environment dependence, whereas we find a significant decline in the SDSS LINER fraction in our overdense bins. The explanation for this difference may come from our removal of possible contaminating star forming galaxies by restricting our analysis to the red sequence. Also, we impose a higher line detection threshold on the SDSS data to provide a fair comparison with DEEP2 (see Section~\\ref{sec:sdss}). Finally, \\cite{Kauffmann2003} required that all lines for the BPT diagnosis were detected and this implies biasing the LINERs sample towards the strongest objects. Between redshifts $z=0.4$ and $z=1.35$, \\cite{Cooper2007a} show that \\emph{red galaxies} within the DEEP2 survey favour overdense environments, although the blue fraction in clusters does become larger as one moves to higher redshift \\citep[see also][]{Gerke2007, Coil2007b}. At all redshifts there exists a non-negligible red fraction in underdense environments, which evolves only weakly if at all. \\cite{Nandra2007} show that the host DEEP2 galaxies of X-ray selected AGN within the EGS field ($\\sim 1/6$ of the DEEP2 survey volume) occupy a unique region of colour-magnitude space. These objects typically live at the top of the blue cloud, within the green valley, or on the red sequence. \\cite{Georgakakis2007} measure the mean environment of this population and confirm that, on average, they live in density regions above that of the mean of the survey. They find this to be true for all host galaxy magnitudes studied ($M_B\\simlt-21$) and colours ($U-B\\simgt0.8$) (note the DEEP2 red sequence begins at $U-B\\sim1$). However, given limited sample sizes, they were not able to establish whether the environment distribution of the X-ray AGN differed from that of the red population, rather than the DEEP2 population as a whole. \\subsection{Understanding the sequence of events} From our results alone a comprehensive understanding of the different environment trends within the AGN population from high to low redshift is not possible. However, some speculation and interpretation can be made by drawing on our broader knowledge of these active objects from the literature. One possible scenario posits that LINERs and Seyferts occur in different types of galaxies. In this picture, LINERs are often associated with young red sequence galaxies \\citep{Graves2007} and are especially common among post-starburst (K+A) galaxies \\citep{Yan2006}. These galaxies would already be into the quenched phase of their evolution but still relatively young. Merger triggered starbursts and subsequent quasar winds are a possible mechanisms to produce rapid star formation shut down in such objects \\citep{Hopkins2006}. The gas rich merging events required in this scenario are common in overdense environments at $z\\sim1$ as clusters and massive groups assemble. By $z\\sim0$, however, the activity in these environments has mostly ended. Hence, if this picture is correct, one may expect an over-abundance of red sequence LINERs in dense environments at high redshift (since both star formation and rapid quenching is common) that is not seen locally. This may be consistent with the trends found in the left panel of Figure~\\ref{fig:SDSS_result}. Seyfert galaxies, on the other hand, could be objects in transition from the blue cloud to red sequence \\citep{Groves2006}, whose AGN are thought to be initiated by internal processes (and not mergers), inferred from their often found spiral structure (e.g. M77) (mergers act to destroy such structure). From this, one may expect our red sequence Seyfert population to represent the tail of the colour distribution of transitioning objects whose dependence on environment is determined by secular mechanisms and who would evolve accordingly. At high redshift, disk galaxies are commonly found in all environments, including the most dense. In contrast, overdense regions in the local Universe are dominated by passive ellipticals and show an absence of spirals. This would be broadly consistent with our findings in the right panel of Figure~\\ref{fig:SDSS_result}, where the most significant evolution in the red Seyfert fraction arises from a depletion in overdense regions relative to other environments, from high redshift to low. Alternatively, some authors claim that LINERs and Seyferts form a continuous sequence, with the Eddington rate the primary distinguishing factor \\citep{Kewley2006}. In this scenario, Seyferts are young objects with actively accreting black holes. As the star formation begins to decay so does the accretion rate, and the galaxy enters a transition phase. Eventually, a LINER-like object emerges, with an old stellar population and very low supermassive BH accretion rate. This picture is supported by recent studies in voids from \\cite{Constantin2007}. At high redshift, our results show that red Seyferts and LINERs are approximately equally abundant. By $z\\sim0$ however, the Seyfert population has declined relative to the LINER population by over a factor of $7$. This may be interpreted as the natural transformation of Seyferts into LINERs with time, within a galaxy population which is smoothly reddening from $z\\sim1$ to $z\\sim0$. Moreover, the fact that high-z LINERs reside preferentially in high-density environments may imply that this Seyfert-LINER transition is more efficient in dense regions of the Universe. In this paper we measure the dependence of the AGN fraction of red galaxies on environment in the $z\\sim1$ DEEP2 Galaxy Redshift Survey and local $z\\sim0.1$ SDSS. We restrict our analysis to the red sequence to maintain a clean and consistent selection of AGN at high and low redshift, and this also reduces the additional effects of environment associated with galaxy colour. Our results can be summarised as follows: \\begin{itemize} \\item[(i)] High redshift LINERs at $z\\sim1$ in DEEP2 appear to favour higher density environments relative to the red sequence from which they are drawn. In contrast, Seyferts and X-ray selected AGN at $z\\sim1$ show much weaker (or no) environmental dependencies within the same underlying population. Extending our analysis to include green valley objects has little effect on the results. \\item[(ii)] Low redshift LINER and Seyfert AGN in the SDSS both show a slowly declining red sequence AGN fraction towards high density environments. This is in contrast to the high redshift result. \\item[(iii)] At $z\\sim1$, Seyferts and LINERs are approximately equally abundant. By $z\\sim0$ however, the Seyfert population has declined relative to the LINER population by over a factor of $7$. \\end{itemize} It is important to remember that such measures are difficult to make with current data, and hence we remain limited by statistics to the extent to which we can physically interpret our results. Regardless, a robust outcome of our analysis is the differences between LINER and Seyfert AGN populations in high density regions, and between high and low redshift in all environments. Our results indicate that a greater understanding of both AGN and galaxy evolution may be possible if future analyses simultaneously focus on the detailed subdivision of different AGN classes, host galaxy properties, and their environment." }, "0801/0801.2902_arXiv.txt": { "abstract": "The coronal magnetic field is an important quantity because the magnetic field dominates the structure of the solar corona. Unfortunately direct measurements of coronal magnetic fields are usually not available. The photospheric magnetic field is measured routinely with vector magnetographs. These photospheric measurements are extrapolated into the solar corona. The extrapolated coronal magnetic field depends on assumptions regarding the coronal plasma, e.g. force-freeness. Force-free means that all non-magnetic forces like pressure gradients and gravity are neglected. This approach is well justified in the solar corona due to the low plasma beta. One has to take care, however, about ambiguities, noise and non-magnetic forces in the photosphere, where the magnetic field vector is measured. Here we review different numerical methods for a nonlinear force-free coronal magnetic field extrapolation: Grad-Rubin codes, upward integration method, MHD-relaxation, optimization and the boundary element approach. We briefly discuss the main features of the different methods and concentrate mainly on recently developed new codes. ", "introduction": "$ $\\\\ Information regarding the coronal magnetic field are important for space weather application like the onset of flares and coronal mass ejections (CMEs). Unfortunately we usually cannot measure the coronal magnetic field directly, although recently some progress has been made \\citep[see e.g.,][]{judge98,solanki:etal03,lin:etal04}. Due to the optically thin coronal plasma direct measurements of the coronal magnetic field have a line-of-sight integrated character and to derive the accurate 3D structure of the coronal magnetic field a vector tomographic inversion is required. Corresponding feasibility studies based on coronal Zeeman and Hanle effect measurements have recently been done by \\cite{kramar:etal06} and \\cite{kramar:etal07}. These direct measurements are only available for a few individual cases and usually one has to extrapolate the coronal magnetic field from photospheric magnetic measurements. To do so, one has to make assumptions regarding the coronal plasma. It is helpful that the low solar corona is strongly dominated by the coronal magnetic field and the magnetic pressure is orders of magnitudes higher than the plasma pressure. The quotient of plasma pressure $p$ and magnetic pressure, $B^2/(2 \\mu_0)$ is small compared to unity $(\\beta=2 \\mu_0 p/B^2 \\ll 1)$. In lowest order non-magnetic forces like pressure gradient and gravity can be neglected which leads to the force-free assumption. Force free fields are characterized by the equations: \\begin{eqnarray} {\\bf j}\\times{\\bf B} & = & {\\bf 0} \\label{forcebal}\\\\ \\nabla \\times {\\bf B }& = & \\mu_0 {\\bf j} \\label{ampere}, \\\\ \\nabla\\cdot{\\bf B} & = & 0 \\label{solenoidal}, \\end{eqnarray} where ${\\bf B}$ is the magnetic field, ${\\bf j}$ the electric current density and $\\mu_0$ the permeability of vacuum. Equation (\\ref{forcebal}) implies that for force-free fields the current density and the magnetic field are parallel, i.e. \\begin{equation} \\mu_0 {\\bf j} = \\alpha {\\bf B}, \\label{jparb} \\end{equation} or by replacing ${\\bf j}$ with Eq. (\\ref{ampere}) \\begin{equation} \\nabla \\times {\\bf B } = \\alpha {\\bf B} \\label{amperealpha}, \\end{equation} where $\\alpha$ is called the force-free function. To get some insights in the structure of the space dependent function $\\alpha$, we take the divergence of Eq. (\\ref{jparb}) and make use of Eqs. (\\ref{ampere}) and (\\ref{solenoidal}): \\begin{equation} {\\bf B} \\cdot \\nabla \\alpha = 0, \\label{alphaeq} \\end{equation} which tells us that the force-free function $\\alpha$ is constant on every field line, but will usually change from one field line to another. This generic case is called nonlinear force-free approach. Popular simplifications are $\\alpha=0$ (current free potential fields, see e.g., \\cite{schmidt64,semel67,schatten69,sakurai82}) and $\\alpha={\\rm constant}$ (linear force-free approach, see e.g., \\cite{nakagawa:etal72,chiu:etal77,seehafer78,alissandrakis81,seehafer82,semel88}). These simplified models have been in particular popular due to their relative mathematical simplicity and because only line-of-sight photospheric magnetic field measurements are required. Linear force-free fields still contain one free global parameter $\\alpha$, which can be derived by comparing coronal images with projections of magnetic field lines (e.g., \\cite{carcedo:etal03}). It is also possible to derive an averaged value of $\\alpha$ from transverse photospheric magnetic field measurements (e.g. \\cite{pevtsov:etal94,wheatland99,leka:etal99}). Despite the popularity and frequent use of these simplified models in the past, there are several limitations in these models (see below) which ask for considering the more sophisticated nonlinear force-free approach. Our aim is to review recent developments of the extrapolation of nonlinear force-free fields (NLFFF). For earlier reviews on force-free fields we refer to \\citep{sakurai89,aly89,amari:etal97,mcclymont:etal97} and chapter 5 of \\cite{aschwanden:book}. Here we will concentrate mainly on new developments which took place after these earlier reviews. Our main emphasis is to study methods which extrapolate the coronal magnetic field from photospheric vector magnetograms. Several vector magnetographs are currently operating or planed for the nearest future, e.g., ground based: the solar flare telescope/NAOJ \\citep{sakurai:etal95}, the imaging vector magnetograph/MEES Observatory \\citep{mickey:etal96}, Big Bear Solar Observatory, Infrared Polarimeter VTT, SOLIS/NSO \\citep{henney06} and space born: Hinode/SOT \\citep{shimizu04}, SDO/HMI \\citep{borrero:etal06}. Measurements from these vector magnetograms will provide us eventually with the magnetic field vector on the photosphere, say $B_{z0}$ for the normal and $B_{x0}$ and $B_{y0}$ for the transverse field. Deriving these quantities from the measurements is an involved physical process, which includes measurements based on the Zeeman and Hanle effect, the inversion of Stokes profiles (e.g., \\cite{labonte:etal99}) and removing the $180^\\circ$ ambiguity (e.g., \\cite{metcalf94,metcalf:etal06}) of the horizontal magnetic field component. Special care has to be taken for vector magnetograph measurements which are not close to the solar disk, when the line-of-sight and normal magnetic field component are far apart (e.g. \\cite{gary:etal90}). For the purpose of this paper we do not address the observational methods and recent developments and problems related to deriving the photospheric magnetic field vector. We rather will concentrate on how to use the photospheric $B_{x0}, \\, B_{y0}$ and $B_{z0}$ to derive the coronal magnetic field. The transverse photospheric magnetic field $(B_{x0}, \\, B_{y0})$ can be used to approximate the normal electric current distribution by \\begin{equation} \\mu_0 j_{z0}=\\frac{\\partial B_{y0}}{\\partial x}-\\frac{\\partial B_{x0}}{\\partial y} \\label{j_photo} \\end{equation} and from this one gets the distribution of $\\alpha$ on the photosphere by \\begin{equation} \\alpha(x,y)=\\mu_0 \\frac{j_{z0}}{B_{z0}} \\label{alpha0direct} \\end{equation} By using Eq. (\\ref{alpha0direct}) one has to keep in mind that rather large uncertainties in the transverse field component and numerical derivations used in (\\ref{j_photo}) can cumulate in significant errors for the current density. The problem becomes even more severe by using (\\ref{alpha0direct}) to compute $\\alpha$ in regions with a low normal magnetic field strength $B_{z0}$. Special care has to be taken at photospheric polarity inversion lines, i.e. lines along which $B_{z0}=0$ \\citep[see e.g.,][]{cuperman:etal91}. The nonlinear force-free coronal magnetic field extrapolation is a boundary value problem. As we will see later some of the NLFFF-codes make use of (\\ref{alpha0direct}) to specify the boundary conditions while other methods use the photospheric magnetic field vector more directly to extrapolate the field into the corona. Pure mathematical investigations of the nonlinear force-free equations \\citep[see e.g.][]{aly84,boulmezaoud:etal00,aly05} and modelling approaches not based on vector magnetograms are important and occasionally mentioned in this paper. A detailed review of these topics is well outside the scope of this paper, however. Some of the model-approaches not based on vector magnetograms are occasionally used to test the nonlinear force-free extrapolation codes described here. \\subsection{Why do we need nonlinear force-free fields?} \\begin{itemize} \\item A comparison of global potential magnetic field models with TRACE-images by \\cite{schrijver:etal05a} revealed that significant nonpotentially occurs regulary in active regions, in particular when new flux has emerged in or close to the regions. \\item Usually $\\alpha$ changes in space, even inside one active region. This can be seen, if we try to fit for the optimal linear force-free parameter $\\alpha$ by comparing field lines with coronal plasma structures. An example can be seen in \\cite{wiegelmann:etal02} where stereoscopic reconstructed loops by \\cite{aschwanden:etal99} have been compared with a linear force-free field model. The optimal value of $\\alpha$ changes even sign within the investigated active regions, which is a contradiction to the $\\alpha={\\rm constant}$ linear force-free approach (see Fig. \\ref{figure1}). \\item Photospheric $\\alpha$ distributions derived from vector magnetic field measurements by Eq. (\\ref{alpha0direct}) show as well that $\\alpha$ usually changes within an active region \\citep[see, e.g.][]{regnier:etal02}. \\item Potential and linear force-free fields are too simple to estimate the free magnetic energy and magnetic topology accurately. The magnetic energy of linear force-free fields is unbounded in a halfspace \\citep{seehafer78} which makes this approach unsuitable for energy approximations of the coronal magnetic field. Potential fields have a minimum energy for an observed line-of-sight photospheric magnetic field. An estimate of the excess of energy a configuration has above that of a potential field is an important quantity which might help to understand the onset of flares and coronal mass ejections. \\item A direct comparison of measured fields in a newly developed active region by \\cite{solanki:etal03} with extrapolations from the photosphere with a potential, linear and nonlinear force-free model by \\cite{wiegelmann:etal05} showed that nonlinear fields are more accurate than simpler models. Fig. \\ref{figure2} shows some selected magnetic field lines for the original measured field and extrapolations from the photosphere with the help of a potential, linear and nonlinear force-free model. \\end{itemize} These points tell us that nonlinear force-free modelling is required for an accurate reconstruction of the coronal magnetic field. Simpler models have been used frequently in the past. Global potential fields provide some information of the coronal magnetic field structure already, e.g. the location of coronal holes. The generic case of force-free coronal magnetic field models are nonlinear force-free fields, however. Under generic we understand that $\\alpha$ can (and usually will) change in space, but this approach also includes the special cases $\\alpha={\\rm constant}$ and $ \\alpha=0$. Some active regions just happen to be more potential (or linear force-free) and if this is the case they can be described with simpler models. Linear force-free models might provide a rough estimate of the true 3D magnetic field structure if the nonlinearity is weak. The use of simpler models was often justified due to limited observational data, in particular if only the line-of-sight photospheric magnetic field has been measured. While the assumption of nonlinear force-free fields is well accepted for the coronal magnetic fields in active regions, this is not true for the photosphere. The photospheric plasma is a finite $\\beta$ plasma and nonmagnetic forces like pressure gradient and gravity cannot be neglected here. As a result electric currents have a component perpendicular to the magnetic field, which contradicts the force-free assumption. We will discuss later, how these difficulties can be overcome. ", "conclusions": "Within the last few years the scientific community showed a growing interest into coronal magnetic fields. \\footnote{Publications containing the phrase 'coronal magnetic fields' in title or abstract have been cited less than about $50$ times per year until the early 1990th and this number increased to about $150$ citations per year in 2004. A peak year was 2006 (last year) with more than $300$ citations. (Source: ISI Web of Knowledge, march 2007)} The development of new ground based and space born vector magnetographs provide us measurements of the magnetic field vector on the suns photosphere. Accompanied from these hardware development, software has been developed to extrapolate the photospheric measurements into the corona. Special attention has recently been given to nonlinear force-free codes. Five different numerical approaches (Grad-Rubin, upward integration, MHD-relaxation, optimization, boundary elements) have been developed for this aim. It is remarkable that new codes or major updates of existing codes have been published for all five methods within the last two years, mainly in the last year (2006). A workshop series (NLFFF-consortium) since 2004 on nonlinear force-free fields has recently released synergy effects, by bringing modelers of the different numerical implementations together to compare, evaluate and improve the programs. Several of the most recent new codes and utility programs (e.g. preprocessing) have at least been partly inspired by these workshops. The new implementations have been tested with the smooth semi-analytic Low-Lou-equilibrium and showed reasonable agreement with this reference field. While all methods aim for a reconstruction of the coronal magnetic field from the photospheric magnetic field vector, the way how these measurements are used to prescribe the boundaries of the codes is different. \\begin{itemize} \\item MHD-relaxation and optimization use $B_{x0}, B_{y0}, B_{z0}$ on the bottom boundary. This over-determines the boundary value problem. Both methods are closely related and compute the magnetic field in a computational box with \\begin{equation} \\frac{\\partial {\\bf B}}{\\partial t}= \\mu {\\bf F}, \\end{equation} where the structure of {\\bf F} is somewhat different (the optimization approach has more terms) for both methods. Usually a potential field is used as initial state for both approaches, also the use of a linear force-free initial state is possible. Recently a multiscale version of optimization has been installed, which uses a low resolution NLFFF-field as input for higher resolution computations. Specifying the entire magnetic field vector on the bottom boundary is an over-imposed problem and a unique NLFFF-field (or a solution at all) requires that the boundary data fulfill certain consistency criteria. A recently developed preprocessing-routine helps to find suitable consistent boundary data from inconsistent photospheric measurements. Earlier and current comparisons showed a somewhat higher accuracy for the optimization approach. A practical advantage of the MHD-approach is that in principle any available time-dependent MHD-code can be adjusted to compute the NLFFF-field. \\item The Grad-Rubin approach uses $B_{z0}$ and the distribution of $\\alpha$ computed with Eq. (\\ref{alpha0direct}) for one polarity, which corresponds to well posed mathematical problem. A practical problem is that the computation of $\\alpha$ requires numerical differences of the noisy and forced transverse photospheric field $B_{x0}, B_{y0}$ with (\\ref{j_photo}) leading to inaccuracies in the normal electric current distribution and in $\\alpha$. For smooth semi-analytic test cases this is certainly not a problem, but real data require special attention (smoothing, preprocessing, limiting $\\alpha \\not=0$ to regions where $B_{z0}$ is above a certain limit) to derive a meaningful distribution of $\\alpha$. While the method requires only $\\alpha$ for one polarity, the computation from photospheric data provide $\\alpha$ for both polarities. We are not aware of any tests on how well NLFFF-solutions computed from $\\alpha$ prescribed on the positive and negative polarity coincide. It is also unclear how well the computed transverse field components on the bottom boundary agree with the measured values \\footnote{In principle $B_{x0}, B_{y0}$ may have an additional field $(B_{x0}, B_{y0})+(\\partial_x \\partial_y) \\phi$ without making a difference for $\\alpha$ and hence for the Grad-Rubin result.} of $B_{x0}, B_{y0}$. More tests on this topics are necessary, including the recently installed possibility to prescribe $\\alpha$ for both polarities and adjust the boundary by a weighed average of $\\alpha$ on both polarities to fulfill Eq. (\\ref{alphaeq}). As initial state the Grad-Rubin method uses a potential field, which is also true for MHD-relaxation and optimization. \\item The upward integration and the boundary element method prescribe both all components of the bottom boundary magnetic field vector and the $\\alpha$ distribution computed with Eq. (\\ref{alpha0direct}). This approach over imposes the boundary and $B_{x0}, B_{y0}, B_{z0}$ and $\\alpha$ have to be consistent which each other and the force-free assumption. This is certainly not a problem at all for smooth semi-analytic test equilibria and strategies to derive consistent boundary data from measured data have been developed recently. Different from the three approaches discussed above, upward integration and boundary element methods do not require to compute first an initial potential field in the computational domain. It is well known that the upward integration method is based on an ill-posed problem and the method has not been considered for several years, but a recent implementation with smooth analytic functions might help to regularize this method. First tests showed a reasonable results for computations with the smooth semi-analytic Low-Lou solution. The boundary element method has the problem to be very slow and an earlier implementation of this method could not reach a converged state for a $64^3$ boxed used in the I. NLFFF-consortium paper due to this problem. A new 'direct boundary method' has been developed, which seems to be faster than the original 'boundary element method', but still slower compared with the four other NLFFF-approaches if the task is to compute a 3D magnetic field in an entire 3D-domain. Different from all other described methods the boundary element approach allows to compute the nonlinear force-free field vector at any arbitrary point above the boundary and it is not necessary to compute the entire 3D-field above the photosphere. This might be a very useful feature if one is interested in computing the magnetic field only along a single loop and not interested in an entire active region. The new implementations of upward integration and boundary element method show both reasonable results for first tests with the smooth semi-analytic Low and Lou equilibrium. Further tests with more sophisticated equilibria, e.g. a solar-like test case as used in the II. NLFFF-consortium paper would be useful to come to more sound conclusions regarding the feasibility of these methods. \\end{itemize} Most of the efforts done in nonlinear force-free modelling until now concentrated mainly on developing these models and testing their accuracy and speed with the help of well known test configuration. Not too many applications of nonlinear force-free models to real data are currently available, from which we learned new physics. One reason was the insufficient access to high accuracy photospheric vector magnetograms and a second one were limitations of the models. Force-free field extrapolation is a mere tool, if properly employed on vector magnetograms, it can help to understand physical, magnetic field dominated processes in the corona. Both the computational methods as well as the accuracy of required measurements (e.g. with Hinode, SDO) are rapidly improving. Within the NLFFF-consortium we just started (since april 2007) to apply the different codes to compute nonlinear force-free coronal magnetic fields from Hinode vectormagnetograms. This project might provide us already some new insights about coronal physics. To conclude, we can say that the capability of Cartesian nonlinear force-free extrapolation codes has rapidly increased in recent years. Only three years ago most codes run usually on grids of about $64^3$ pixel. Recently developed or updated codes (Grad-Rubin by Wheatland, MHD-relaxation by Valori, optimization by Wiegelmann, optimization by McTiernan) have been applied to grids of about $300^3$ pixel. Although this increase of traceable grid sizes is certainly encouraging, the resolution of current and near future vector magnetographs (which of course measure only data in 2D!) is significantly higher. We should keep in mind, however, that the currently implemented NLFFF-codes have been only moderately parallelized using only a few processors. The CSWM-conference, where this paper has been presented, took place at the 'Earth simulator' in Yokohama, which contains several thousands of processors used for Earth-science computer simulations. An installation of NLFFF-codes on such massive parallel computers (which has been briefly addressed on NLFFF-consortium meetings) combined with adaptive mesh refinements might enable drastically improved grid sizes. One should not underestimate the time and effort necessary to program and install such massive parallelized versions of existing codes. As full disk vectormagnetograms will become available soon (SOLIS, SDO/HMI) it is also an important task to take a spherical geometry into account. First steps in this direction have been carried out with the optimization and boundary element methods. Spherical NLFFF-geometries are currently still in it's infancy and have been tested until now only with smooth semi-analytic Low and Lou equilibria and require further developments. Attention has also recently been drawn to the problem that the coronal magnetic field is force-free, but the photospheric one is not. Tests with extrapolations from solar-like artificial photospheric and chromospheric measurements within the II. NLFFF-consortium paper revealed that extrapolations from the (force-free) chromospheric field provide significantly better results as extrapolations using directly the (forced) photospheric field. Applying a pre-processing program on the photospheric data, which effectively removes the non-magnetic forces, leads to significantly better results, but they are not as good as by using the chromospheric magnetic field vector as boundary condition. An area of current research is the possibility to use chromospheric images to improve the preprocessing of photospheric magnetic field measurements. Improvements in measuring the chromospheric magnetic field directly \\citep[e.g.][]{lagg:etal04} might further improve to find suitable boundary conditions for NLFFF-extrapolations. Force-free extrapolations are not suitable, however, to understand the details of physical processes on how the magnetic field evolves from the forced photosphere into the chromosphere, because non-magnetic forces are important in the photosphere. For a better understanding of these phenomena more sophisticated models which take pressure gradients and gravity (and maybe also plasma flow) into account are required. Some first steps have been done with a generalization of the optimization method by \\cite{wiegelmann:etal06b}, but such approaches are still in their infancy and have been tested so far only with smooth MHD-equilibria. It is also not entirely clear how well necessary information regarding the plasma (density, pressure, temperature, flow) can be derived from measurements. Non-magnetic forces become important also in quiet sun regions (\\cite{schrijver:etal05}) and in the higher layers of the corona, where the plasma $\\beta$ is of the order of unity. Coronagraph measurements, preferably from two viewpoints as provided by the STEREO-mission, combined with a tomographic inversion might help here to get insights in the required 3D structure of the plasma density. One should also pay attention to the combination of extrapolation methods, as described here, with measurements of the Hanle and Zeeman effects in coronal lines which allows the reconstruction of the coronal magnetic field as proposed in feasibility studies of vector tomography by \\cite{kramar:etal06,kramar:etal07}. Other measurements of coronal features, e.g., coronal plasma images from two STEREO-viewpoints, can be used for observational tests of coronal magnetic field models. Using two viewpoints provide a much more restrictive test of models as images from only one view direction. While a nonlinear force-free coronal magnetic field model helps us to derive the topology, magnetic field and electric current strength in coronal loops, they do not provide plasma parameters. One way to get insights regarding the coronal plasma is the use of scaling laws to model the plasma along the reconstructed 3D field lines and compare correspondent artificial plasma images with real coronal images. \\cite{schrijver:etal04} applied such an approach to global potential coronal magnetic fields and compared simulated and real coronal images from one viewpoint. A generalization of such methods towards the use of more sophisticated magnetic field models and coronal images from two STEREO-viewpoints will probably provide many insights regarding the structure and physics of the coronal plasma. An important challenge is for example the coronal heating problem. The dominating coronal magnetic field is assumed to play an important role here, because magnetic field configuration containing free energy can under certain circumstances reconnect (\\cite{priest96,priest99}) and supply energy for coronal heating. \\cite{priest:etal05} pointed out that magnetic reconnection at separators and separatrices plays an important role for coronal heating. Nonlinear force-free models can help here to identify the magnetic field topology, magnetic null points, separatrices and localized strong current concentration. While magnetic reconnection \\citep[see e.g.][]{priest:etal99} is a dynamical phenomenon, the static magnetic field models discussed here can help to identify the locations favourable for reconnection. Time sequences of nonlinear force-free models computed from corresponding vector magnetograms will also tell wether the topology of the coronal magnetic field has changed due to reconnection, even if the physics of reconnection is not described by force-free models. Sophisticated 3D coronal magnetic field models and plasma images from two viewpoints might help to constrain the coronal heating function further, which has been done so far with plasma images from one viewpoint \\citep[][by using data from Yokoh, Soho and Trace]{aschwanden01,aschwanden01a}." }, "0801/0801.0907_arXiv.txt": { "abstract": "We study dust accumulation by photophoresis in optically thin gas disks. Using formulae of the photophoretic force that are applicable for the free molecular regime and for the slip-flow regime, we calculate dust accumulation distances as a function of the particle size. It is found that photophoresis pushes particles (smaller than 10 cm) outward. For a Sun-like star, these particles are transported to $0.1-100$ AU, depending on the particle size, and forms an inner disk. Radiation pressure pushes out small particles ($\\la 1$ mm) further and forms an extended outer disk. Consequently, an inner hole opens inside $\\sim 0.1$ AU. The radius of the inner hole is determined by the condition that the mean free path of the gas molecules equals the maximum size of the particles that photophoresis effectively works on ($100 \\ \\micron - 10$ cm, depending on the dust property). The dust disk structure formed by photophoresis can be distinguished from the structure of gas-free dust disk models, because the particle sizes of the outer disks are larger, and the inner hole radius depends on the gas density. ", "introduction": "Protoplanetary disks are composed of gas and dust. At initial stages of their evolution, the gas of protoplanetary disks is as massive as $10^{-3} - 10^{-1} M_{\\sun}$ (Greaves 2004), and small dust particles are mixed with the gas, making the disks optically thick at optical wavelengths. As the dust particles grow, the number of small particles reduces, and at a certain stage, the disks become optically thin (Tanaka et al. 2005; Dullemond \\& Dominik 2005). The amount of gas also decreases, as the gas dissipates (e.g., Hartmann et al. 1998; Clarke et al. 2001; Takeuchi et al. 2005; Alexander et al. 2006a,b). At late stages of the disk evolution, the disks become gas-free dust disks, which are observed as Vega-type disks. During the transition from protoplanetary disks to Vega-type disks, there may be a stage where the disks are optically thin, but their gas component still remains. An example is HD 141569A, which is a 5 Myr Herbig Be star (Weinberger et al. 2000) and has an optically thin dust disk (Augereau et al. 1999; Weinberger et al. 1999; Fisher et al. 2000; Mouillet et al. 2001; Marsh et al. 2002; Clampin et al. 2003). The gas component of this system has been observed (Zuckerman 1995; Dent et al. 2005; Goto et al. 2006) and its amount is estimated to be $\\la 60 \\ M_{\\earth}$ (Ardila et al. 2005). Dynamics of dust particles in optically thin gas disks is of interest in order to investigate the structure of transitional disks. Krauss \\& Wurm (2005) considered the motion of dust particles in a gas disk that is optically thin at optical wavelengths. A dust particle receives stellar radiation directly, and the radiation pressure pushes the particle outward. In addition to radiation pressure, interaction between the particle and the surrounding gas molecules induces photophoresis, which also pushes the particle outward (see also Wurm \\& Krauss 2006; Krauss et al. 2007). When these outward accelerations act with the gas drag on the particle, the particle drifts outward (Takeuchi \\& Artymowicz 2001). In the outer part of the disk ($\\ga 10$ AU), where the mean free path of the gas molecules is larger than the particle size (of $\\la 1$ m), the photophoretic force is proportional to the gas density. When a particle moves outward to the point where the gas density is no longer high enough to induce a strong photophoretic force, the particle's outward motion stops. Consequently, the dust particles pile-up at a certain distance from the star that is determined by the density profile of the gas disk. In a gas disk with mass $\\sim 0.01 \\ M_{\\sun}$, particles of $100 \\ \\micron$ to 10 cm pile-up at several tens of AU. This spontaneous ring formation is a characteristic feature that is caused by photophoresis in a gaseous dust disk. Krauss et al. (2007) have demonstrated that photophoretic dust motion may be the key process for the transition from optically thick protoplanetary disks to optically thin circumstellar disks with ring-like dust distributions via the stage of transitional disks with an inner hole and a more or less sharp transition to the outer disk that is continuously pushed outward. This outward dust migration can also explain the presence of material formed close to the sun like chondrules and CAIs in asteroids of the main asteroid belt (Wurm \\& Krauss 2006) or high temperature crystalline silicates in comets from the Kuiper belt (Petit et al. 2006; Brownlee et al. 2006; Mousis et al. 2007). In this paper, we seek to investigate other characteristic structures formed by photophoresis, especially in the inner part of the disk, but at a stage when most of the dust has already been built into larger bodies. Krauss \\& Wurm (2005) focus on the dust dynamics in the outer part of the disk ($\\ga 10$ \\ AU) and use a formula for the photophoretic force that is relevant for low gas densities. In order to investigate the structure of the whole disk, we use a photophoresis formula that can be applied to the whole range of gas densities. Details of the formula for the photophoretic force adopted here are described in \\S\\ref{sec:drift}. Using this formula, we find that photophoresis does not induce dust pile-up at the innermost region of the disk ($\\la 0.1$ AU), leading to the formation of an inner hole (see \\S\\ref{sec:accrad}). At the inner part of the gas disk, the gas density may be too high that the gas disk itself is no longer optically thin to the stellar radiation, and thus any radiation effects on the dust such as radiation pressure and photophoresis do not work at all. In order to investigate whether photophoresis still works at the inner disk, in \\S\\ref{sec:opacity}, we calculate the optical depth of the disk due to Rayleigh scattering of the hydrogen molecules. In \\S\\ref{sec:discussion1}, the thermal relaxation time of dust particles is estimated and is compared to the rotation periods induced by gas turbulence and by photophoresis. In \\S\\ref{sec:discussion2}, we discuss what are the characteristic features that photophoresis forms in dust disks, and what are the observable differences from the gas-free dust disk structures. In Appendix A, a simple calculation deriving the photophoretic force is described. ", "conclusions": "\\subsection{Thermal Relaxation and Rotation Times of Dust Particles} \\label{sec:discussion1} As mentioned in \\S\\ref{sec:fdust}, if particles rotate rapidly, photophoresis does not work. Even if particle rotation is not considerably rapid compared to the thermal relaxation time of the particle, photophoretic Yarkovsky effect may prevent the particle accumulation. Such cases where photophoresis does not work are represented by the model of $J_1=0$, where the incident starlight does not cause any temperature gradient in the particles. If photophoresis is suppressed by any reason, particles can reside in the region where ${\\rm Kn} < 1$ (Fig. \\ref{fig:req}), and such cases can be observationally distinguished from the disks in which photophoresis clears the dust in the innermost region. Hence, whether particles in gas disks rotate rapidly or not can be observationally investigated by checking whether the dust in the region of ${\\rm Kn} < 1$ is cleared or not. In this subsection, we estimate the thermal relaxation time of a dust particle and then compare it to the rotation period induced by gas turbulence or by photophoresis itself. We use a simple model of a cylindrical dust particle described in Appendix A. Consider a cylindrical dust particle with radius $a$ and height $2a$ (see Figure \\ref{fig:cylndust}). The stellar radiation flux $I$ irradiates the front surface. We suppose that, at the beginning, the temperature inside the particle is homogeneous. This initial (or average) temperature $T_d$ is given by equation (\\ref{eq:td}), assuming that the incident flux $I$ balances with thermal radiation from the front and back surfaces. (Radiation from the side surface is ignored for simplicity.) The temperature of the front surface increases to the equilibrium value $T_f$ in a thermal relaxation time $\\tau_{\\rm th}$. During $\\tau_{\\rm th}$, the incident energy at the front surface conducts for a length $\\delta a$ and forms a skin layer of temperature gradient. During $\\tau_{\\rm th}$, the temperature at the front surface is approximated as the initial value $T_d$, and the energy flux inside the skin layer is estimated as $Q_{\\rm con} \\approx I-\\sigma_{\\rm SB} T_d^4 = I/2$. (We set $\\varepsilon=1$.) When the skin layer depth has grown to $\\delta a$ and the temperature of the front surface converges to $T_f$, the energy flux is $Q_{\\rm con} \\approx k_d (T_f - T_d) / \\delta a$. Equating the above two expressions for $Q_{\\rm con}$, the thickness of the temperature gradient layer is \\begin{equation} \\delta a = \\frac{2 k_d T_d \\Delta}{I} \\ , \\end{equation} where $\\Delta = (T_f - T_d)/T_d$. Note that the maximum value of $\\delta a$ is $2a$. The thermal relaxation time is \\begin{equation} \\tau_{\\rm th} = \\frac{c_d \\rho_d \\delta a^2}{k_d} \\ , \\end{equation} where $c_d$ is the specific heat capacity of the particle. \\begin{figure} \\epsscale{1.2} \\plotone{f7.eps} \\caption{ Thermal relaxation times of cylindrical particles of various sizes are plotted by the solid lines against the distance from the star. The dashed line shows the turnover time, $\\tau_{\\rm ed}$, of the smallest eddies for $\\alpha_{\\rm tur}=10^{-2}$. \\label{fig:tauth} } \\end{figure} Figure \\ref{fig:tauth} shows the thermal relaxation time $\\tau_{\\rm th}$ for $10 \\ \\micron - 10$ cm particles in the disk of model A. We set the particle bulk density $\\rho_d=1 \\ {\\rm g \\ cm^{-3}}$, the specific heat capacity $c_d=10^7 \\ {\\rm erg \\ g^{-1} \\ K^{-1}}$, and the thermal conductivity $k_d=10^2 \\ {\\rm erg \\ s^{-1} \\ cm^{-1} \\ K^{-1}}$. The temperature difference $\\Delta$ is calculated by equation (\\ref{eq:tempdif}). At small distances from the star, radiative cooling $4 \\varepsilon \\sigma_{\\rm SB} T_d^4$ dominates in determining $\\Delta$ in equation (\\ref{eq:tempdif}), and $\\Delta \\approx 1/4$. Thus, $\\tau_{\\rm th}$ is independent of the particle size $a$, and is proportional to $T_d^{-6}$. In our model $T_d \\propto r^{-1/2}$, and thus $\\tau_{\\rm th} \\propto r^3$. At large distances, on the other hand, internal thermal conduction $k_d T_d /a$ determines $\\Delta$. In this case, the skin depth is $\\delta a=a$. Hence, $\\tau_{\\rm th}$ is proportional to $a^2$ and independent of $r$. If the rotation period of the particle is smaller than $\\tau_{\\rm th}$, photophoresis is considerably suppressed. Particle rotation can be excited by several mechanisms such as Brownian motion, gas turbulent motion, collisions with other dust particles, and the photophoretic force itself can induce rotation if it has a offset from the mass center. Detailed calculation of the rotation period for each mechanism is beyond the scope of this paper. Here, we make a rough estimate of the rotation speed induced by turbulence and by the photophoretic force. Suppose that the gas disk has isotropic turbulence with the Kolmogorov energy spectrum. We assume that the largest eddies have size $L=\\alpha_{\\rm tur}^{1/2} h_g$ and velocity $V=\\alpha_{\\rm tur}^{1/2} c_s$, where $\\alpha_{\\rm tur}$ is the ``$\\alpha$-viscosity'' parameter (Cuzzi et al. 2001). The energy of turbulent motion cascades down to smaller eddies, and finally dissipates by the molecular viscosity. From dimensional analysis, the energy dissipation rate per unit mass is $\\dot{e} \\sim V^3 / L \\sim \\alpha_{\\rm tur} c_s^2 \\Omega_{\\rm K}$. The eddy turnover time is faster for smaller eddies and that for the smallest eddies is \\begin{equation} \\tau_{\\rm ed} \\sim \\left( \\frac{\\eta_{\\rm vis}}{\\rho_g \\dot{e}} \\right)^{1/2} \\sim \\alpha_{\\rm tur}^{-1/2} \\left( \\frac{l}{h_g} \\right)^{1/2} \\Omega_{\\rm K}^{-1} \\ , \\end{equation} where $\\eta_{\\rm vis}=l v_T \\rho_g /2$ is the molecular viscosity (Weidenschilling 1984). The smallest eddies can induce a particle rotation period as short as $\\tau_{\\rm ed}$, if the particle is well coupled to the turbulent motion of the smallest eddies. However, if the particle does not strongly couple to the gas, the rotation period is longer than $\\tau_{\\rm ed}$. In Figure \\ref{fig:tauth}, the turnover time, $\\tau_{\\rm ed}$, of the smallest eddies is plotted by the dashed line for $\\alpha_{\\rm tur} = 10^{-2}$. The particle rotation period induced by gas turbulence is expected to be larger than the dashed line, and therefore is much longer than the thermal relaxation time. \\begin{figure} \\epsscale{1.2} \\plotone{f8a.eps} \\plotone{f8b.eps} \\caption{ Times taken for a particle to rotate 180 degree by photophoresis are plotted by the dashed lines. The photophoretic force is exerted on a point at a distance $b=a$ or $b=0.01a$ from the cylinder axis. The solid line shows the thermal relaxation time. ($a$) For $100 \\ \\micron$ particles. ($b$) For 1 cm particles. \\label{fig:rot} } \\end{figure} We next consider rotation induced by the torque exerted by photophoresis. Consider again a cylindrical dust particle and suppose that the photophoretic force is exerted on a point at a distance $b$ from the cylinder axis. The torque is $K=b F_{\\rm ph}$. The value of $b$ is unknown, and we treat $b$ as a free parameter. The principal moment of inertia of the cylinder (for the axis perpendicular to the cylinder axis) is $I_{xx}=7 \\pi a^5 \\rho_d / 6$. The time needed for a initially stationary particle to rotate 180 degree is \\begin{equation} \\tau_{\\rm rot} = \\left( \\frac{2 \\pi I_{xx}}{K} \\right)^{1/2} = \\left( \\frac{7 \\pi^2 \\rho_d a^5}{3 b F_{\\rm ph}} \\right)^{1/2} \\ . \\end{equation} In Figure \\ref{fig:rot}, the rotation time $\\tau_{\\rm rot}$ is plotted for assumed off-centers $b=a$ and $b=0.01a$. If the photophoretic force has a large off-center ($b=a$), the rotation time can be shorter than the thermal relaxation time at $0.1 - 2$ AU for $100 \\ \\micron$ particles and at $4-40$ AU for 1 cm particles. In such regions, photophoresis is probably significantly suppressed. In order for photophoresis to work effectively in the whole disk, the off-center of the photophoretic force has to be as small as $b=0.01a$. \\subsection{Characteristic Structure Due to Photophoresis} \\label{sec:discussion2} The dust disk structure that photophoresis makes has three zones: the outer disk, the inner disk, and the inner hole. The outer disk is composed of small ($a \\la 1$ mm or $\\beta \\ga 0.01$) particles, and their dynamics is controlled mainly by radiation pressure. The inner disk is composed of large ($a \\ga 100 \\ \\micron$ or $\\beta \\la 0.01$) particles accumulating there due to photophoresis. The boundary between the outer and inner regions is at $10-100$ AU, depending on the gas density profile. The inner hole opens inside $\\sim 0.1$ AU, where the Knudsen number ${\\rm Kn}$ is smaller than unity for particles that photophoresis effectively works on ($a=100 \\ \\micron - 10$ cm). The structure formed by photophoresis should be compared to the structure of gas-free disks. Modeling of gas-free dust disks also shows zonal structure that consists of the outer extended disk, the inner disk, and the inner hole (Th\\'ebault \\& Augereau 2005; Wyatt 2006; Strubbe \\& Chiang 2006; Krivov et al. 2006). In their models, a planetesimal belt is assumed and dust particles are continuously produced by planetesimal collisions. The inner disk is composed of large particles whose orbits are hardly affected by radiation pressure and are nearly circular. Thus, the location and width of the inner disk are basically similar to those of the planetesimal belt. Outside the planetesimal belt, the outer disk extends and it is composed of small particles whose $\\beta$-value is large. Their orbits are strongly influenced by radiation pressure and are excited to highly eccentric orbits. In the model of Th\\'ebault \\& Augereau (2005), the particles of the outer disk have $\\beta > 0.05$, which is relatively larger than $\\beta$ of our models. In our models, particles' $\\beta$ of the outer disk can be as small as $0.01$. Therefore, infrared to millimeter radio observations of outer disks and determination of the particle size will provide key information for determining which model is plausible. Our models of photophoresis also predict that an inner hole opens in the dust disk, but the gas still fills the hole. This is a significant feature, but it is difficult to test this with the present observational techniques. We have to observe the gas, but the amount of the gas of optically thin disks is probably quite small, and most of Vega-type stars do not have gas in an amount exceeding the current detection limit. Further, the radius of the inner hole is of the order of $0.1$ AU, and the temperature there is close to the sublimation temperature of the dust. (The temperature exceeds 1500 K inside $0.03$ AU of our model disks.) Therefore, we need a careful observation that can distinguish between the inner holes made by photophoresis and by dust sublimation. \\subsection{Summary} Dust accumulation by photophoresis is studied. We use formulae of the photophoretic force that are applicable for the free molecular regime and for the slip-flow regime. The main results are as follows. 1. Particle accumulation occurs at a point where the outward acceleration on the gas by the pressure gradient equals to the outward acceleration on the particle by radiation pressure and photophoresis. 2. Photophoresis makes an inner disk composed of relatively large particles ($a= 100 \\ \\micron - 10$ cm). The inner disk extends between $0.1$ AU and $10-100$ AU, surrounded by the outer disk composed of small particles ($a \\la 1$ mm). 3. An inner hole opens inside $\\sim 0.1$ AU. The inner hole radius is determined by the condition ${\\rm Kn}=1$ for the maximum size particles that photophoresis effectively works on ($a=100 \\ \\micron - 10$ cm). Photophoresis works effectively only when the disk is optically thin. Most of small ($\\la 10 \\ \\micron$) dust grains must be removed from the disk such that their column density to the star becomes smaller than $10^{-3} \\ {\\rm g \\ cm^{-2}}$. For example, at 1 AU, the dust density must be smaller than $10^{-16} \\ {\\rm g \\ cm^{-3}}$, i.e., $10^{-5}$ times smaller than the value of the minimum mass solar nebula model (Hayashi et al. 1985). The gas disk also must be optically thin. Figure \\ref{fig:gasden} shows that the gas surface density must be smaller than $10^2 - 10^3 \\ {\\rm g \\ cm^{-2}}$. At $0.1$ AU, where the typical ray from the star enters the disk (for the dust particles at $r \\gg 0.1$ AU), this value is $\\sim 10^{-2}$ times smaller than the value of the minimum mass solar nebula model (Hayashi et al. 1985). Even in such a tenuous gas disk, the photophoretic force is strong enough to change the dust disk structure inside a few AU, as shown by model A (Figure \\ref{fig:req}$a$). If the gas density is more tenuous than model A, the region where photophoresis has a substantial effect shrinks toward the central star. In a gas disk with $0.1$ times smaller gas density than model A, photophoresis on millimeter sized particles works effectively only inside 1 AU, and in a disk with $10^{-2}$ times smaller gas density, the effective region shrinks to $0.3$ AU." }, "0801/0801.2133_arXiv.txt": { "abstract": "I calculate the action of a satellite, infalling through dynamical friction, on a coplanar gaseous disk of finite radial extent. The disk tides, raised by the infalling satellite, couple the satellite and disk. Dynamical friction acting on the satellite then shrinks the radius of the coupled satellite-disk system. Thus, the gas is ``shepherded'' to smaller radii. In addition, gas shepherding produces a large surface density enhancement at the disk edge. If the disk edge then becomes gravitationally unstable and fragments, it may give rise to enhanced star formation. On the other hand, if the satellite is sufficiently massive and dense, the gas may be transported from $\\sim 100$ pc to inside of a 10 to 10s of parsecs before completely fragmenting into stars. I argue that gas shepherding may drive the fueling of active galaxies and central starbursts and I compare this scenario to competing scenarios. I argue that sufficiently large and dense super star clusters (acting as the shepherding satellites) can shepherd a gas disk down to ten to tens of parsecs. Inside of ten to tens of parsecs, another mechanism may operate, i.e., cloud-cloud collisions or a marginally (gravitationally) stable disk, that drives the gas $\\lesssim 1$ pc, where it can be viscously accreted, feeding a central engine. ", "introduction": "\\label{sec:intro} The tidal interaction between small satellites and the gas or particle disks in which they are embedded is a subject of wide study in the planetary community. These satellites are tidally coupled via disk tides, i.e., satellites excite spiral density waves at the Lindblad resonances in the disk (Goldreich \\& Tremaine 1978, 1980; Artymowicz 1993). As a result of these tidal interactions, gaps can be opened up in the embedded disks as in the case of the shepherding moons of Saturn's rings (Goldreich \\& Tremaine 1978) or in type-II migration (Ward 1997). Alternatively, if a gap does not open up, satellite may migrate inward rapidly due to tidal torque imbalances, i.e., type-I migration (Ward 1997). The wide applicability of satellite-disk interactions in the planetary community raises an interesting question of whether the same physics may be applicable at larger scales, i.e., galactic scales. In this paper, I will address this question by considering a very simple model, which illustrate the modification of the physics of satellite-disk interaction when applied on a galactic scale. Most notably, the critical difference is that in addition to tidal torques, the satellite will experience dynamical friction. The inclusion of dynamical friction produces a non-trivial effect. Namely, dynamical friction on the satellite provides a sink of angular momentum in the system. As a result, the coupled satellite-disk system will continually lose angular momentum and sink toward the center. To study this physics, I consider a very simple model. In my model, a satellite starts out in a circular orbit at a large radius and sinks toward the central mass concentration because of dynamical friction on the background stars. Along its infall, it encounters a coplanar gaseous disk or ring, which initially has a finite radial extent, $r_{\\rm d,0}$. Tides begin to couple the satellite with the disk. Because the satellite-disk system continues to suffer an ongoing loss of angular momentum from dynamical friction, it will shrink in radius. As a result gas can be transported on the dynamical friction timescale to smaller radii. In addition, I find that this satellite-disk interaction also builds a substantial surface density enhancement at the disk edge. This may lead to enhanced star formation at the disk edge. This process which I call {\\it gas shepherding} may be a generic feature of gaseous disks around galaxies, if a sufficiently massive and dense satellite is available. The physics of satellite-disk interactions in galaxies or gas shepherding is not just an interesting exercise in mathematical physics, but may be important in the fueling of active galaxies and central starbursts. First, the action of dynamical friction on the satellite-disk system will shrink radius of the disk, thereby transporting gas to smaller radii. This shepherding of gas will continue until the gas is forced into the center or the shepherding satellite is destroyed. I will discuss this scenario for nuclear fueling and how it compares to competing scenarios in this paper. I have structured this paper as follows. In \\S\\ref{sec:basic picture}, I discuss the basic picture of gas shepherding and give a few order of magnitude estimates. I estimate the timescale for gas shepherding and the expected surface density enhancement. I present the physics of gas shepherding and solve numerical models in \\S\\ref{sec:shepherding}. I also present an approximate analytic solution to this problem. I then discuss the application of gas shepherding to the feeding of active galactic nuclei and central starbursts in \\S\\ref{sec:applications}. Finally I present my conclusions in \\S\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} I have calculated the action of an infalling satellite on a coplanar gaseous disk. The satellite falls inward because of dynamical friction and couples to the gaseous disk via disk tides. Due to the unrelenting loss of angular momentum, the satellite-disk system shrinks in radius. Thus, gas is transported from larger radii to smaller radii. I calculate the structure of the evolving disk as a function of radius, both numerically and analytically and show a significant density enhancement at the disk edge. This density enhancement is due to the transport of gas initially at larger radii to a ring that is $r_{\\rm H,d}$ thick radially. The fate of this shepherded gas is not known. I presented two possible outcomes: 1. the gas will fragment and form stars or 2. the gas is shepherded into the nuclear region. In the latter case, I suggest that the gas shepherding scenario, which I have presented may be one means by which the ``100 parsec'' problem may be resolved. For this to occur, gas must be shepherded sufficiently rapidly such that the gas is not completely consumed in star formation, which requires a sufficiently massive satellite ($\\sim 10^6\\,{\\rm M}_{\\odot}$). I argue that such massive satellites exist on the form of SSCs. In the model presented, the physics of star formation is a disk environment has not been fully addressed. Observationally, galactic star-forming disks appear to be marginally unstable to fragmentation, i.e., $Q \\sim 1$ (Thompson et al 2005) and their star formation rate follows the Kennicutt law (Kennicutt 1998). These two points suggests that some sort of feedback is responsible for maintaining this careful balance between fragmentation and heating. How this feedback operates is not well understood. As discussed previously, various models have invoked radiation pressure (Thompson et al. 2005), supernova feedback (Wada \\& Norman 2002), and cosmic ray pressure (Socrates et al. 2008) from star formation. Alternatively, the turbulent pressure support may result from gravitational instability, i.e., gravitoturbulence (Gammie 2001). It is unclear that in the presence of gas shepherding if this equilibrium can be maintained at the edge of the disk where the surface density is enhanced. If it does not, this edge is likely to fragment completely into stars on a few dynamical times. It is also unclear if the Kennicutt law would continue to hold at the disk edge. Additional insights from future research on self-gravitating star-forming galactic disks will be needed to help formulate a more self-consistent calculation of the gas shepherding scenario, which will be needed to fully address the fate of the shepherded gas. Assuming that the gas is not fully consumed in star formation, it would be shepherd down to 10 to 10s of parsecs (see \\S \\ref{sec:final 10 pc}). At that radii cloud-cloud collisions (Krolik \\& Begelman 1988; SBF90) or a starburst supported disk (Thompson et al. 2005) would allow the material to be viscously accreted. The initial satellite inclination and/or eccentricity is likely to significantly affect gas shepherding. In this paper, I have studied a special case where the orbit of the satellite is in the plane of the disk and approximately circular for simplicity. An eccentric satellite would have its eccentricity damped via the excitation of spiral density waves. If the eccentricity is below $r_{\\rm H, d}/r_{\\rm d}$, i.e., the satellite stays within one ``Hill radius'' of the disk of its outer edge, then the shepherding calculation presented should be a reasonable estimate of the shepherding timescales and velocities. Compared to a coplanar satellite, an inclined satellite would couple less well to the gaseous disk. For a sufficiently large inclination, the gas disk and satellite would be effectively decoupled. However, its inclination could be sufficiently rapidly damped via induced bending waves in the gas disk in addition to spiral density waves (see for instance Ostriker 1994; Terquem 1998). Once its inclination is damped below the critical value, where the disk and satellite become well coupled, the shepherding scenario presented in this paper would likely follow. I would expect this critical inclination to be of order $\\sin i = r_{\\rm H,d}/r_{\\rm d}$, i.e., the satellite stays within of order a ``Hill radius'' of the disk above and below it. Extending the present calculation to satellites of modest inclination would be fruitful and would allow a study of using the nucleated cores from minor mergers as shepherding satellites (see \\S\\ref{sec:shepherd destruction}). As discussed in \\S\\ref{sec:applications}, minor mergers may drive nuclear activity. The initial inclination of minor mergers is arbitary, but their inclination is damped as they interact with the {\\it large scale} stellar and gaseous disk. As a result, they may approach the shepherding scenario presented in this paper. Such a study is a subject for future work. In addition, dynamical friction would also act on the satellite-induced spiral density waves. Dynamical friction of the extended mass perturbation, i.e., the satellite-induced spiral density waves, may enhance the transfer of angular momentum between the gas disk and background stars (Tremaine \\& Weinberg 1984). A study of these effects, while fruitful, is beyond the scope of this work. The physics of gas shepherding is also important in other areas as well. For instance, it may shape radial distribution of star clusters in nuclear gas rings. In some of these nuclear ring systems, the location of the star clusters is exterior to the gas ring from which they are presumably formed. For instance, NGC 4314 clearly show the star clusters exterior to the gas (Benedict et al. 2002). Martini et al. (2003) found that the eight galaxies with strong bars all have star formation exterior to the dust ring in their sample of 123 galaxies. How gas shepherding drives this morphology is the subject of a forthcoming work. (Van der Ven \\& Chang, in preparation)." }, "0801/0801.3724.txt": { "abstract": "% % Context: ------------------- { Rings or annulus-like features have been observed in most imaged debris discs. Outside the main ring, while some systems (e.g., $\\beta$ Pictoris and AU Mic) exhibit smooth surface brightness profiles (SB) that fall off roughly as $\\sim r^{-3.5}$, others (e.g. HR 4796A and HD 139664) display large drops in luminosity at the ring's outer edge and steeper radial luminosity profiles. } % % Aims: ------------------------ { We seek to understand this diversity of outer edge profiles under the ``natural'' collisional evolution of the system, without invoking external agents such as planets or gas. } % % Methods: { We use a multi-annulus statistical code to follow the evolution of a collisional population, ranging in size from dust grains to planetesimals and initially confined within a belt (the \"birth ring\"). The crucial effect of radiation pressure on the dynamics and spatial distribution of the smallest grains is taken into account. We explore the dependence of the resulting disc surface brightness profile on various parameters. } % % Results: { The disc typically evolves toward a ``standard'' steady state, where the radial surface brightness profile smoothly decreases with radius as $r^{-3.5}$ outside the birth ring. This confirms and extends the semi-analytical study of Strubbe \\& Chiang (2006) and provides a firm basis for interpreting observed discs. Deviations from this typical profile, in the form of a sharp outer edge and a steeper fall-off, occur for two \"extreme\" cases: 1) When the birth ring is so massive that it becomes radially optically thick for the smallest grains. However, the required disc mass is probably too high here to be realistic. 2) When the dynamical excitation of the dust-producing planetesimals is so low ($$ and $ \\leq 0.01$) that the smallest grains, which otherwise dominate the optical depth of the system, are preferentially depleted. This low-excitation case, although possibly not generic, cannot be ruled out by observations for most systems, . } %", "introduction": "\\subsection{the ubiquity of ring-like features} Dusty debris discs have been detected by their infrared excess around $\\sim 15\\%$ of nearby main sequence stars \\citep[e.g.][]{back93}. More than a dozen of these discs have also been imaged, mainly in scattered light, since the initial observation of the $\\beta$ Pictoris system by \\citet{smi84}. One unexpected result from these images is that almost no system displays a smooth extended radial profile: the usual morphology is the presence of rings (or annuli) where the bulk of the dust population is located. This ring morphology is in fact so common that \\citet{stru06} pointed out that the debris disc phenomenon could more appropriately be renamed debris ``ring'' phenomenon. Even for the archetypal debris ``disc'' $\\beta$ Pictoris, which has been imaged from 5 to a few thousand AU, the bulk of the dust is probably concentrated in a rather narrow region between 80 and 120 AU \\citep[e.g.][]{aug01}. One of the few systems actually resembling an extended ``smooth'' disc might be Vega, for which Spitzer mid--infrared images show a rather smooth radial luminosity profile \\citep{su05}. At the other end of disc morphologies, among the most striking ring features are the ones around HR4796 \\citep[e.g][]{Jay98,koer98,schnei99}, HD139664 \\citep{kal06}, and Fomalhaut \\citep{kal05}. \\begin{table*} \\begin{minipage}{\\textwidth} \\caption[]{Geometry and surface brightness profile for a selection of debris disc systems resolved in scattered light images. } \\label{init} \\renewcommand{\\footnoterule}{} \\begin{tabular}{lcccc} System \\footnote{For many of these systems, many additional features, i.e., warps, clumps, etc., have been observed (it is especially true for \\bp) but we focus here on the main issue of average radial profiles}& Orientation & Detected Radial Extent\\footnote{The radial extents and surface brightness profiles are given for regions beyond the likely ``birth ring'' -- the region where scattering luminosity peaks and where most parent bodies are believed to reside (see Sec.\\,\\ref{sec:approach})} & Surface Brightness $SB \\propto r^{\\alpha}$ & Reference\\\\ \\hline HD 53143 & $\\sim$ face-on & $55-110$AU & $\\alpha \\sim -3$ & Kalas et al. (2006)\\\\ $\\beta$ Pic & edge-on & $127-193$AU & $\\alpha \\simeq -3$ & Golimovski et al.(2006) \\\\ & & $193-258$AU& $\\alpha \\simeq -4$ & \\\\ HD 32297 & edge-on & $80-170$AU & $\\alpha \\sim -3.5$ & \\citet{schnei99}\\\\ & & $170-350$AU & $\\alpha \\sim -3$ (averaged over the 2 ansae)& \\\\ AU Mic & edge-on & $32-210$AU & $\\alpha \\sim -3.8$ (averaged over the 2 ansae) & Fitzgerald et al (2007) \\\\ HD139664 & edge-on & $83-109$AU& $\\alpha \\sim -4.5$ & Kalas et. al. (2006)\\\\ HD 107146 & $\\sim$ face-on & $130-185$AU & $\\alpha \\sim -4.8 \\pm 0.3$ & Ardila et al. (2004)\\\\ HD 181327 & face on & $100-200$ AU & $\\alpha \\sim -5$ & \\citet{schnei06}\\\\ Fomalhaut & $66^{o}$ inclination & $140-160$ AU & $\\alpha \\sim -6.1$ & Kalas et al. (2006)\\\\ HR 4796A & $73^{o}$ inclination & $70-120$AU & $\\alpha \\sim -7.5$ & \\citet{wahhaj05} \\\\ \\hline \\end{tabular} \\label{table:observed} \\end{minipage} \\end{table*} The characteristics that most differentiates one ring-like system from the other is the sharpness of the luminosity drop at the inner and outer edges of the rings. We shall in this paper focus on the outer edge issue, for which systems can be basically divided into two categories \\citep[see also][]{kal06}: \\begin{itemize} \\item ``Sharp edge'' rings, displaying abrupt surface brightness drops, sometimes as steep as $r^{-8}$ beyond the ring. The most representative members of this group are HR4796A and Fomalhaut. \\item ``Smooth edge'' rings, with no sharp outer edge and a surface brightness drop beyond the ring in $r^{-3}$ or $r^{-4}$. The most famous examples are here $\\beta$ Pic and AU Mic \\end{itemize} See Tab.\\ref{table:observed} for a list of outer-edge profiles for several resolved debris discs. The presence of ring features is commonly attributed to some sculpting mechanisms including, in particular, the presence of a massive planet (see, for instance, \\citet{quillen07} for Fomalhaut, \\citet{frey07} for $\\beta$ Pictoris, \\citet{wyatt99} for HR4796A, or the more general study of \\citet{moro05}). The gravitational effect of such a planet can truncate or create gaps in the disc, either directly on the dust particles or indirectly on the planetesimals that produce these particles. \\subsection{outer edges} Nevertheless, while planets may be a natural and easy explanation for sculpting the inner edges, outer edges are a different problem. They are difficult to explain in the light of one well established fact about debris discs, i.e. that the observed dust is not primordial but steadily produced by collisions, through a collisional cascade starting at much larger parent bodies, maybe in the planetesimal size range \\citep[e.g.][]{lag00}. In this respect, even if there is a sharp outer edge for the parent body population, collisions would constantly produce very small grains that would be launched by radiation pressure onto eccentric or even unbound orbits, thus populating the region beyond the outer edge and erasing the appearance of a narrow ring over timescales which might be shorter than those for gravitational sculpting by a planet. This issue is a critical one, since these small grains dominate the total geometric cross section and thus the flux in scattered light \\citep[see the discussion in][]{theb07}. One possible confining mechanism for the sharp outer edge is gas. For discs transiting between gas-dominated phase (proto-planetary discs) to dust-dominated phase (debris discs), narrow dust rings may arise \\citep{klahr05}. \\citet{beslawu} further demonstrate that there exists an instability with which the residual gas collects grains of various sizes (even those subject to radiation pressure) into a narrow belt. However, this mechanism requires at least comparable amount of gas and dust. This may be difficult to justify for most debris discs, which are evolved systems where the amount of gas is probably too low to prevent the smallest grains to be launched onto very eccentric orbits smoothing out any sharp outer edge. ``Razor sharp'' outer edges are thus very difficult to explain in the presence of this unavoidable outward launching of small grains. However, although no perfect abrupt outer edge has indeed been observed \\footnote{HD141569A could be a possible exception, but this system is likely not a true debris disc, being significantly younger and gas rich \\citep{jonk06}, but a member of the loosely defined \"transition object\" category.} (unlike inner edges, which are in some cases, like Fomalhaut, almost razor sharp), a great variety of outer edge profiles does exist, from relatively smooth to very steep (see Tab.\\,\\ref{table:observed}). In this study, we address the issue of how these different profiles can be physically achieved: can this diversity be explained by the sole ``natural'' evolution of a collisional active disc steadily producing small, radiation-pressure affected grains, or is (are) additional mechanism(s) needed? We consider initial conditions which are {\\it a priori} the most favourable for creating sharp edges, by assuming a population of large parent bodies confined within an annulus with an abrupt cutoff at its outer edge. How this initial confinement may have come about is itself an interesting question but is not the focus of the current paper (see however the discussion in Sec.\\,\\ref{sec:real}). The outcome we consider as a reference for our investigation is the radial surface brightness (SB) profile in scattered light, since this is an observable which is reasonably well constrained for most imaged debris discs (either directly observed or obtained by de-projection). We consider the nominal case of a disc seen edge-on, but results can easily be extrapolated to face-on systems, since SB profiles for both orientations tend towards the same radial dependence far from the birth ring (given the same radial dust distribution). In several previous studies, all addressing the specific $\\beta$ Pic case \\citep{lec96, aug01, theb05}, it has been argued that the ``natural'' SB profile outside the collisionaly active parent bodies belt, or ``birth ring'', falls off as $\\sim r^{-5}$. This is based on the assumption that all particles produced in the birth ring have a size distribution which scales as $dN/ds \\propto s^{-3.5}$, \\citep[as expected for an idealized infinite collisional cascade at equilibrium, see][]{dohn69}, down to the radiation blow--out limit $s=s_{0.5}$ (where the ratio of radiation pressure to gravity $\\beta = 0.5$). The smallest radiation pressure-affected grains, which dominate the light receiving area, are then diluted along their eccentric orbit. This geometrical spread results in $SB \\propto r^{-5}$. However, \\citet{stru06} argued that, since high-$\\beta$ grains spend a long time in the collisionaly inactive region beyond the birth ring, the $dN/ds \\propto s^{-3.5}$ collisional equilibrium law should only apply to the small fraction $f$ of these grains which are present in the collisionaly active birth ring. This results in a $1/f$ excess of the disc-integrated number of small grains, which in turn results in a flatter $SB$ profile in $r^{-3.5}$. They applied their theory to the AU Mic disc \\footnote{AU Mic is an M star with weak radiation but where stellar wind is believed to act on small grains in an equivalent way as radiation pressure does around more massive A stars \\citep[e.g][]{aug06}.} and reproduced the observed SB profile, spectral energy distribution and disc colour. \\citet{stru06} further argued that the observed SB profile depends only weakly on the radial and size distributions of grains within the birth ring. The discs which exhibit a fall-off sharper than $r^{-3.5}$ are thus puzzling in the face of this theory. The innovative model of \\citet{stru06} is build on analytical derivations and Monte-Carlo modeling which did not actually treat the collisional evolution of the system and relies on several simplifying assumptions. The main one is that the size distribution is fixed and is assumed to follow the idealized $dN/ds \\propto s^{-3.5}$ scaling (corrected by the fraction $1/f$), whereas several studies have shown that this law cannot hold in real systems \\citep[see][and references therein]{theb07} \\footnote{The need for realistic size distributions departing from fixed power laws has been very recently emphasized by \\citet{fitz07} in their analysis of the AUMic data}. Another issue is that when evaluating collisional life-times, only the vertical velocity of the grains was taken into account, thus neglecting their radial movement which can be appreciable, if not dominant for the smallest grains. Finally, the specific dynamics of the small radiation-pressure-affected grains, in particular the fact that they suffer much more frequent collisions and at much higher velocities, is not taken into account. ", "conclusions": "We numerically investigate if the diverse outer surface brightness profiles observed in debris discs can be explained by the ``natural'' collisional evolution of belts made of solid particles ranging from planetesimal to micron-sized dust grains. We consider an initial ring of parent bodies with a razor-sharp outer edge and quantitatively examine to what extent the steady collisional production of small, radiation-pressure-affected grains modifies the initially perfect ring-like structure. Concurring with the pioneering results of \\citet{stru06}, our numerical explorations have shown that for most ``reasonable'' parameters (mass, ring width, dynamical excitation) of a collisionally evolving debris ring, the surface brightness outside the ring naturally tends towards a standard profile, with no sharp drop at the outer edge and a mid-plane surface brightness profile $\\propto r^{-3.5}$. We confirm this result by simple but robust analytical considerations. This nominal result is in good agreement with the luminosity profiles observed in the outer regions of several debris discs, including \\bp\\ and AUMic. Some observed systems however exhibit steeper radial luminosity profiles. Our numerical exploration shows that sharper outer edges and profiles steeper than that of the nominal case can only be obtained for two ``extreme'' cases: \\begin{itemize} \\item 1) For a system with very high dust content (typically $M_{dust}\\geq10M_{\\oplus}$), the radial optical depth is raised to near unity for small grains. Most of these high-$\\beta$ grains cannot travel out of the birth ring without suffering a collision. This allows the outer edge to sharpen. However, systems with such high radial optical depth are expected to wear down with time because of strong collisional erosion. Moreover, the existence of such very dusty systems is not backed by observations. This case does not appear as a realistic option. \\item 2) A longer survival of the sharp outer edge is achieved for systems with normal dust content that are \\emph{dynamically cold}, with typically $\\langle e \\rangle =2\\langle i \\rangle \\leq 0.01$. In this case, small grains are destroyed much more efficiently than they are created, leading to a depletion of this population. The system's optical depth and luminosity are then dominated by large grains which do not leave the main birth ring, leading to a sharp outer edge. Even if this case might not correspond to the most generic debris disc configuration, it cannot be ruled out by observations and is thus a possible explanation to some of the observed systems. \\end{itemize} To numerically investigate the applicability of the dynamically cold case to real sharp-edge systems, we consider the specific case of HR4796A. We find a reasonably good fit of this system's outer region luminosity profile with a dynamical excitation $\\langle e \\rangle \\sim 0.0035$. There is thus the possibility that such a sharp outer edge could be explained by the natural collisional evolution of a confined disc of large parent bodies. %(the confinement of the parent bodies %being possibly due to some past punctual or transitory event or %to the natural outer limit of the protoplanetary nebulae)." }, "0801/0801.3952_arXiv.txt": { "abstract": "In 1981, a faint radio source (G$'$) was detected near the center of the lensing galaxy of the famous ``twin quasar'' Q0957+561. It is still unknown whether this central radio source is a third quasar image or an active nucleus of the lensing galaxy, or a combination of both. In an attempt to resolve this ambiguity, we observed Q0957+561 at radio wavelengths of 13~cm, 18~cm, and 21~cm, using the Very Long Baseline Array in combination with the phased Very Large Array and the Green Bank Telescope. We measured the spectrum of G$'$ for the first time and found it to be significantly different from the spectra of the two bright quasar images. This finding suggests that the central component is primarily or entirely emission from the foreground lens galaxy, but the spectrum is also consistent with the hypothesis of a central quasar image suffering free-free absorption. In addition, we confirm the previously-reported VLBI position of G$'$ just north of the optical center of the lens galaxy. The position slightly favors the hypothesis that G$'$ originates in the lens, but is not conclusive. We discuss the prospects for further clarification of this issue. ", "introduction": "When a galaxy lies close to the line of sight to a quasar, multiple images of the quasar may be produced. Almost all known lensed quasars have an even number of images---either two or four---despite a mathematical proof that nonsingular lens models always produce an odd number of images \\citep{dyer80a,burke81a}. The missing image is the ``central image,'' corresponding to the central maximum in light-travel time. Central images are hard to detect because they are highly demagnified by the dense core of the lensing galaxy. They are worth finding, however, because the properties of central images are unique probes of the inner 10-100 parsecs of galaxies that are too distant to resolve with ordinary observations. Even non-detections of central images can be useful constraints on the central structure of galaxies. An upper limit on the flux density of a central image corresponds to a lower bound on the central surface density of the lens galaxy. In the context of particular galaxy models, the absence of a detectable central image can be used to set the maximum size of any constant-density core \\citep{narasimha86a,wallington93a} or the steepness of any central density cusp \\citep{rusin01a,evans02a,keeton03b}. Supermassive black holes also affect central images: they can prevent the formation of a central image, or produce an additional image \\citep{mao01a,bowman04a,rusin05a}. Microlensing by stars in the lens galaxy can demagnify a central image even further \\citep{bernstein99a,dobler07a}. Besides the faintness of the central image, there are two other reasons why it has proven difficult to identify central images. One reason is the close proximity between the expected position of the central image and the center of the lensing galaxy. The expected separation is less than $\\sim$10 milliarcseconds for an isolated lensing galaxy, although it can be larger when the galaxy is part of a cluster that provides additional magnification \\citep{inada05a}. Resolving such a small separation would often require space-based imaging at optical or near-infrared wavelengths, or Very Long Baseline Interferometry (VLBI) at radio wavelengths. A second and related reason is that even when a central point source is detected, it is difficult to decide conclusively whether it is a central image of a background quasar, or an active galactic nucleus (AGN) in the lens, or a combination of these two possibilities. This decision is difficult not only because the point source is faint and challenging to characterize, but also because propagation effects may alter the properties of the central image and cause it to appear different from the other lensed images, thereby interfering with the usual tests for lensed image identification. Such propagation effects include extinction by dust at optical wavelengths, and scintillation and free-free absorption at radio wavelengths; these effects are expected to be more important for lines of sight passing very near the center of the foreground galaxy, where on average the density of dust and plasma should be greatest. A central image was detected for the bright optical lens APM~08279+5255 \\citep{ibata99a,egami00a,munoz01a}, but it seems probable that this image is the result of a highly elongated mass distribution rather than the high central density of the lensing galaxy \\citep{lewis02a}. More recently, \\cite{winn04a} presented evidence for a central image in PMN~J1632--0033, thereby setting an upper limit on the mass of any central black hole and a lower limit on the central surface density of the lens galaxy. A central image has also been detected at optical wavelengths for SDSS~J1004+4112 \\citep{inada05a}. In addition, there are cases in which central radio components have been detected but are most likely to be the active nucleus of the lens galaxy \\citep{chen93a,fassnacht99b}, and cases where recent work has placed stringent upper limits on the flux density of any central component \\citep{boyce06a,zhang07a}. This paper is concerned with the very first lensed quasar to have been discovered, Q0957+561 \\citep{walsh79a}. More than 20 years ago, a central radio component was detected in this system \\citep{gorenstein83a}, but there is still a lingering uncertainty about its interpretation. The system consists of a quasar at redshift 1.41 that is gravitationally lensed into 2 images (A and B) by an elliptical galaxy G1 and a surrounding set of galaxies mostly at redshift 0.36 \\citep{stockton80a,young81a}. Quasar images A and B appear as point sources in optical images and as core-jet sources in radio images. Observations with the Very Large Array (VLA) reveal a radio source, dubbed G, close to the center of the optically-identified lens galaxy G1 \\citep{greenfield80a,roberts85a}. Radio observations with VLBI have revealed an unresolved point source near the center of G1 \\citep{gorenstein83a}. The VLBI source is known by the different name G$'$ because its relationship to G is unclear. The VLA and VLBI observations are sensitive to very different spatial scales of radio emission; typically, VLBI observations have an angular resolution 10-100 times finer than VLA observations and are blind to most structures that are resolved by the VLA. Modelers of this lens have typically assumed that G$'$ marks the center of the lensing galaxy, and that any central quasar image is undetectably faint \\citep[see, e.g.,][]{barkana99a,bernstein99a,chae99a}. If instead G$'$ is the central image, then the lens galaxy has a shallower central gravitational potential than has been assumed by those modelers. Traditionally this identification has not been a major concern, because the lens modeling literature has been preoccupied with the connection between the Hubble constant and measurements of the A--B time delay, which is not greatly affected by the innermost $\\sim$100~pc of the galaxy mass distribution. However, as noted above, the interpretation of G$'$ does have an impact on our understanding of the mass distribution of the central regions of elliptical galaxies. We undertook new radio observations of Q0957+561 in an attempt to clarify the nature of G$'$. Our strategy was to compare the radio spectral-energy distributions of A, B, and G$'$ over as large a range of radio wavelengths as feasible. For all practical purposes, gravitational lensing is wavelength-independent. Thus, the central image would have the same spectrum as images A and B, were it not for the confounding factors of intrinsic variability (in conjunction with different time delays for the different images), the magnification gradient along the jet images (in conjunction with different spectral energy distributions for the core and jets), and propagation effects specific to the line of sight of each image (such as scintillation and free-free absorption), which must be taken into account. An AGN in the lens would not have the same spectrum as the quasar images except by coincidence. Although the Q0957+561 system has been studied extensively at radio wavelengths, flux-density measurements of G$'$ have not been well documented. Previously, the only flux-density measurement with a reported uncertainty was by \\cite{gorenstein83a} at 13~cm. The VLA source G has been documented more completely, including measurements at wavelengths of 2~cm, 3.6~cm, and 6~cm by \\cite{harvanek97a}. In the following section, we describe our VLBI observations of this system and G$'$ in particular. We describe the data reduction procedures in \\S\\ref{reduction}, and we present the empirical results in \\S\\ref{results}. In \\S\\ref{discussion} we discuss the spectrum, and in \\S\\ref{future} we review the prospects for further progress on this issue. ", "conclusions": "" }, "0801/0801.4936.txt": { "abstract": "Following the recent abundance measurements of Mg, Al, Ca, Fe, and Ni in the black hole X-ray binary \\mbox{XTE J1118+480} using medium-resolution Keck~II/ESI spectra of the secondary star (Gonz\\'alez Hern\\'andez et al. 2006), we perform a detailed abundance analysis including the abundances of Si and Ti. These element abundances, higher than solar, indicate that the black hole in this system formed in a supernova event, whose nucleosynthetic products could pollute the atmosphere of the secondary star, providing clues on the possible formation region of the system, either Galactic halo, thick disk, or thin disk. We explore a grid of explosion models with different He core masses, metallicities, and geometries. Metal-poor models associated with a formation scenario in the Galactic halo provide unacceptable fits to the observed abundances, allowing us to reject a halo origin for this X-ray binary. The thick-disk scenario produces better fits, although they require substantial fallback and very efficient mixing processes between the inner layers of the explosion and the ejecta, making quite unlikely an origin in the thick disk. The best agreement between the model predictions and the observed abundances is obtained for metal-rich progenitor models. In particular, non-spherically symmetric models are able to explain, without strong assumptions of extensive fallback and mixing, the observed abundances. Moreover, asymmetric mass ejection in a supernova explosion could account for the required impulse necessary to launch the system from its formation region in the Galactic thin disk to its current halo orbit. ", "introduction": "The low-mass X-ray binary \\mbox{XTE J1118+480} is the first identified black hole moving in Galactic halo regions (Wagner et al. 2001; Mirabel et al. 2001). Since it was discovered during a faint outburst on UT 2000 March 29 (Remillard et al.\\ 2000), it has been intensively studied in both the X-ray and optical spectral regions. During the decay of the outburst, McClintock et al. (2001) and Wagner et al. (2001) determined the radial-velocity curve of the companion star, yielding a mass function $f(M) \\approx 6$~\\Msuno. The companion star was classified as a late-type main-sequence star with a mass of 0.1--0.5~{\\Msuno} (Wagner et al.\\ 2001). By modelling the light curve, McClintock et al. (2001) derived a lower limit to the orbital inclination, $i \\ga 55^\\circ$, and consequently an upper limit to the black hole mass of $M_{\\rm BH} \\la 10$~\\Msuno. Additional evidence for a high inclination ($i \\ga 60^\\circ$) comes from measurements of tidal distortion (Frontera et al. 2001), whereas the lack of dips or eclipses for a Roche-lobe filling secondary yields upper limits of $i \\ga 80^\\circ$ and $M_{\\rm BH} \\ga 7.1$~\\Msuno. Later, Gelino et al. (2006) derived an orbital inclination of $68^\\circ\\;\\pm\\;2^\\circ$, by modeling the optical and infrared ellipsoidal light curves of the system in quiescence. This value of the inclination allowed them to better constrain the black hole mass at $M_{\\rm BH} = 8.53 \\pm 0.60$~\\Msuno. The system is placed in the Galactic halo, with an extraordinarily high Galactic latitude ($b \\approx 62.3^{\\circ}$), and a height of $\\sim$1.6 kpc above the Galactic plane, according to its distance of $1.85\\pm0.36$ kpc (Wagner et al.\\ 2001). This appears surprising since all other black hole binaries are located in the Galactic disk. An accurate measurement of its proper motion coupled with its distance provides space-velocity components ($U$, $V$) which seem consistent with those of some old halo globular clusters (Mirabel et al.\\ 2001). This opened the possibility that the system originated in the Galactic halo, and therefore, that the black hole could be either the remnant of a supernova (SN) in the very early Galaxy or the result of direct collapse of an ancient massive star. However, the galactocentric orbit crossed the Galactic plane many times in the past, and the system could have formed in the Galactic disk and been launched into its present orbit as a consequence of the ``kick'' imparted during the SN explosion of a massive star (Gualandris et al.\\ 2005). Recent observations with the 10-m Keck~II telescope revealed that the secondary star has a supersolar surface metallicity ($[{\\rm Fe/H}]=0.2\\pm0.2$, Gonz\\'alez H\\'ernandez et al. 2006), confirming the origin of the black hole in a SN event. Thus, if the system originated in the Galactic halo, the element abundances of the secondary star must have been enriched by a factor of 5--25 depending on whether its initial metallicity resembled a thick-disk star or a halo star. Element abundances of secondary stars of X-ray binaries have been studied for the systems Nova Scorpii 1994 (Israelian et al. 1999; Gonz\\'alez Hern\\'andez et al. 2007), A0620--00 (Gonz\\'alez Hern\\'andez et al. 2004), Centaurus X-4 (Gonz\\'alez Hern\\'andez et al. 2005), \\mbox{XTE J1118+480} (Gonz\\'alez Hern\\'andez et al. 2006, hereafter Paper~I), and V4641 Sagittarii (Orosz et al. 2001; Sadakane et al. 2006). All of these X-ray binaries show metallicities close to solar independent of their location with respect to the Galactic plane, and possible scenarios of pollution from a SN or hypernova have been discussed. In this paper, we compare in detail different scenarios of the possible enrichment of the secondary star from SN yields, providing conclusions on the formation region (Galactic halo, thick disk, or thin disk) of this halo black hole X-ray binary. \\begin{figure}[ht!] %\\epsscale{.80} %\\plotone{f1.eps} \\centering \\includegraphics[width=6.2cm,angle=90]{f1.eps} \\caption{\\footnotesize{Radial velocities of \\mbox{XTE J1118+480} folded on the orbital solution of the data, together with the best-fitting sinusoid. Individual velocity uncertainties are $\\le 7$ ${\\rm km}\\ {\\rm s}^{-1}$ and are not plotted because they are always smaller than the symbol size. The bottom panel shows the residuals of the fit.}} \\label{fig1} \\end{figure} ", "conclusions": "We have presented Keck~II/ESI medium-resolution spectroscopy of the black hole binary \\mbox{XTE J1118+480}. The individual spectra of the system allowed us to derive an orbital period of $P = 0.16995 \\pm 0.00012$~d and a radial velocity semiamplitude of the secondary star of $K_2 = 708.8 \\pm 1.4$ \\kmso. The implied updated mass function is $f(M) = 6.27 \\pm 0.04$~\\Msun, consistent with (but more precise than) previous values reported in the literature. Inspection of the high-quality averaged spectrum of the secondary star provides a rotational velocity of $v~\\sin~i = 100^{+3}_{-11}$ \\kmso, and hence a binary mass ratio $q = 0.027 \\pm 0.009$. The derived radial velocity, $\\gamma = 2.7 \\pm 1.1$ \\kmso, of the center of mass of the system agrees, at the 3$\\sigma$ level, with the results of previous studies. We have performed a detailed chemical analysis of the secondary star. We applied a technique that provides a determination of the stellar parameters, taking into account any possible veiling from the accretion disk. We find $T_{\\mathrm{eff}} = 4700 \\pm 100$~K, $\\log [g/{\\rm cm~s}^2] = 4.6 \\pm 0.3$, $\\mathrm{[Fe/H]} = 0.18 \\pm 0.17$, and a disk veiling (defined as $F_{\\rm disk}/F_{\\rm total}$) of $\\sim$40\\% at 5000~{\\AA}, decreasing toward longer wavelengths. We have provided further details on the abundances of Mg, Al, Ca, Fe, Ni, and Li already reported by Gonz\\'alez Hern\\'andez et al. (2006), and we determined new element abundances of Si and Ti. The chemical abundances are typically higher than solar, and in some cases they are slightly enhanced (e.g., Mg, Al, and Si) in comparison with the abundances of these elements in stars of the solar neighborhood having similar iron content. The present location and kinematics of this binary system had suggested that it could have originated in the Galactic halo. However, the chemical abundances strongly indicate that the black hole formed as a consequence of a supernova/hypernova explosion that occurred within the binary system. This explosive event must have either provided a kick to the system if it was formed in the thin disk or enriched significantly the atmosphere of the secondary star if the system formed in the thick disk or halo. We have explored a variety of supernova/hypernova explosion models for different metallicities, He core masses, and geometries. We compared the expected abundances in the secondary star after contamination from nucleosynthetic products from different initial metallicities of the secondary star ($-2.2 < \\mathrm{[Fe/H]} < 0.2$), to investigate the formation region in the Galactic halo, thick disk, or thin disk. Metal-poor explosion models ($Z = 0$ and $Z = 0.001$) were not able to fit the observed abundances since they produce inappropriate ratios between $\\alpha$-elements and iron-peak elements, and they are extremely sensitive to the adopted mass-cut. This comparison probably rules out an origin in the Galactic halo for this black hole binary. For the thick-disk scenario, we carefully inspected the model predictions, and although they provide better fits to the observed abundances, they require substantial fallback (up to 5.5~\\Msun) and very efficient mixing processes between the inner layer of the explosion and the ejecta. We thus conclude that this scenario is unlikely. Metal-rich spherically symmetric models for the thin-disk scenario were able to fairly reproduce the observed abundances, although they do not easily provide the energy required to launch the system from the Galactic plane to its current halo orbit. Finally, non-spherically symmetric models produce excellent agreement with the observed element abundances in the secondary star without invoking extensive fallback and mixing. In addition, asymmetric mass ejection would naturally provide the kick to expel this binary system from its birth place in the Galactic thin disk, which seems to be the most plausible explanation for the origin of this halo black hole X-ray binary." }, "0801/0801.2435_arXiv.txt": { "abstract": "We present a calculation of the sedimentation of grains in a giant gaseous protoplanet such as that resulting from a disk instability of the type envisioned by Boss (1998). Boss (1998) has suggested that such protoplanets would form cores through the settling of small grains. We have tested this suggestion by following the sedimentation of small silicate grains as the protoplanet contracts and evolves. We find that during the course of the initial contraction of the protoplanet, which lasts some $4\\times 10^5$ years, even very small ($>1\\mu $m) silicate grains can sediment to create a core both for convective and non-convective envelopes, although the sedimentation time is substantially longer if the envelope is convective, and grains are allowed to be carried back up into the envelope by convection. Grains composed of organic material will mostly be evaporated before they get to the core region, while water ice grains will be completely evaporated. These results suggest that if giant planets are formed via the gravitational instability mechanism, a small heavy element core can be formed due to sedimentation of grains, but it will be composed almost entirely of refractory material. Including planetesimal capture, we find core masses between 1 and 10 M$_{\\oplus}$, and a total high-Z enhancement of $\\sim$ 40 M$_{\\oplus}$. The refractories in the envelope will be mostly water vapor and organic residuals. ", "introduction": "The mechanism of giant plant formation is still a matter of debate. The two main candidates are core accretion (see, e.g. Pollack et al. 1996) and disk instability (see, e.g. Boss 1997). In the core accretion scenario, kilometer-sized planetesimals collide and accrete to form a core. As this core grows, it begins to attract the surrounding gas. By the time the core has reached a mass of some 10 $M_{\\oplus}$, the accretion rate of the gas becomes very high and a Jupiter-mass object is formed (Pollack et al. 1996, Hubickyi et al. 2005). \\newpage The alternative scenario invokes a local gravitational instability in the gas disk surrounding a young star. Such an instability can lead to the creation of gas clumps which evolve to become giant plants. Such a mechanism of planet formation would seem to imply that the resultant planet has a solar ratio of elements. Observations of Jupiter indicate that the planet's envelope is enriched in heavy elements by a factor of $\\sim 3$ over the solar ratio to hydrogen (Young 2003). In addition, recent models of Jupiter's interior that fit the gravitational moments (Saumon and Guillot, 2004) indicate that Jupiter's core is between 0--6 $M_{\\oplus}$, smaller than earlier estimations. However, the overall enhancement of heavy elements in the planet, according to these models, is of the order of 20 -- 40 $M_{\\oplus}$, or 3 -- 6 times the solar value. \\noindent The core instability model requires a core of the order of $\\sim 10 M_{\\oplus}$ to reach the rapid gas accretion stage, while the disk instability can, in principle, form a giant planet with no core at all. Indeed a simple interpretation of this scenario would seem to require this. Therefore, if Jupiter does turn out to have a substantial core, the disk instability model must provide a mechanism for forming one. Boss (1997, 1998) has suggested that a solid core can be formed by sedimentation of dust grains to the center of the protoplanet before the protoplanet contracts to planetary densities and temperatures. In this paper we investigate this possibility by applying the microphysics of grain coagulation and sedimentation to silicate grains in a Jupiter-mass protoplanet with initial conditions similar to those in a newly formed clump. ", "conclusions": "We present a calculation of grain sedimentation inside a protoplanet formed via the disk instability scenario. During the first $\\sim 10^5$ yr the clump is cold enough to let silicate grains settle to the center and create a heavy element core. This is true for all grain sizes we considered ($a_0\\geq 10^{-4}$ cm). Grains smaller than this will have high enough number densities so that they will grow quickly into the size range we considered. Core formation proceeds whether the envelope is convective or not, although the convective envelope delays core formation substantially for smaller grains. In all cases a core can be formed before the temperatures near the center get high enough to evaporate the silicate grains. If the silicate material that reaches the central region and is incorporated into a core is immune to further dissolution, core masses of some 4 M$_{\\oplus}$ can be obtained. This includes material added to the planet later via planetesimal capture. If the planetesimals are of the order of 100 km radius or larger, they will not be captured until the protoplanet contracts sufficiently. This delay means that the envelope temperatures are higher when this additional mass settles, and not all the grains can reach the core before they are vaporized. For such large planetesimals, we find that the core mass will be only $\\sim$ 1.7 M$_{\\oplus}$. Grains consisting of ice or more volatile material cannot sediment quickly enough to avoid being vaporized before they reach the core region. Grains consisting of CHON depend on the volatility chosen. For hexacosane the grains evaporate before reaching the core at all the sizes we considered. For CHON that has the volatility suggested by Obrec (2004), only very large grains ($\\gtrsim 10 $ cm) can reach the core, and these add at most a few tenths of an Earth mass to the total core mass. The remainder of this high-Z material would remain in the envelope, giving a high-Z mass in the envelope of some 30 M$_{\\oplus}$. Our simulations thus produce both core and envelope high-Z masses that agree very well with the values determined by fitting the gravitational moments of Jupiter (Guillot 1999, Saumon and Guillot 2004). In this work, two very important effects have been neglected. The first is radiative heating by the surrounding medium. As Kovetz et al. (1988) have shown, such radiation tends to be deposited at a point where the optical depth is approximately equal to the ratio of the incoming flux to the outgoing flux. In our case, the isolated planet initially has a photospheric temperature of $\\sim 25$~K. If we take the protoplanet to be at 5 AU, the intervening gas and dust will have a high optical depth in the visible [e.g. Chiang and Goldreich (1997)], and the radiation falling onto the protoplanet will be mostly at the ambient temperature of the surrounding gas. This is model-dependent, but is of the order of 100 K (Lecar et al. 2006) so that the ratio of incoming to outgoing flux can be quite large. This will cause higher temperatures in the interior, and will make it more difficult for grains to survive long enough to sediment to the core. On the other hand, in the early stages the planet will stay extended longer, and this will allow more time for the grains to settle. A second effect is the contraction of the gas near the center by the increased pressure due to the core itself. This too will heat the gas near the core, and will inhibit the grains from settling. This second effect seems more serious, since, in this scenario, the formation of the core itself inhibits further core growth. A proper treatment of this effect requires an equation of state for the core material, and work on this is in progress." }, "0801/0801.2573_arXiv.txt": { "abstract": "{ Variability is a common characteristic of magnetically active stars. Flaring variability is usually interpreted as the observable consequence of transient magnetic reconnection processes happening in the stellar outer atmosphere. Stellar flares have been observed now across 11 decades in wavelength/frequency/energy; such a large span implies that a range of physical processes takes place during such events. Despite the fact that stellar radio flares have long been recognized and studied, key unanswered questions remain. I will highlight what, in my opinion, are some of these questions. I will also describe recent results on stellar flare emissions at radio wavelengths, discussing the nature of coherent and incoherent emissions and the prospects of wide-field radio imaging telescopes for studying such events. } \\FullConference{Bursts, Pulses and Flickering:Wide-field monitoring of the dynamic radio sky\\ June 12-15 2007\\\\ Kerastari, Tripolis, Greece} \\begin{document} ", "introduction": " ", "conclusions": "" }, "0801/0801.0740_arXiv.txt": { "abstract": "\\noindent We investigate a scenario where the recently discovered non-thermal hard X-ray emission from the Ophiuchus cluster originates from inverse Compton scattering of energetic electrons and positrons produced in weakly interacting dark matter pair annihilations. We show that this scenario can account for both the X-ray and the radio emission, provided the average magnetic field is of the order of \\mbox{0.1 $\\mu$G}. We demonstrate that GLAST will conclusively test the dark matter annihilation hypothesis. Depending on the particle dark matter model, GLAST might even detect the monochromatic line produced by dark matter pair annihilation into two photons. ", "introduction": " ", "conclusions": "" }, "0801/0801.3269_arXiv.txt": { "abstract": "We study the effects of substructure in the Galactic halo on direct detection of dark matter, on searches for energetic neutrinos from WIMP annihilation in the Sun and Earth, and on the enhancement in the WIMP annihilation rate in the halo. Our central result is a probability distribution function (PDF) $P(\\rho)$ for the local dark-matter density. This distribution must be taken into account when using null dark-matter searches to constrain the properties of dark-matter candidates. We take two approaches to calculating the PDF. The first is an analytic model that capitalizes on the scale-invariant nature of the structure--formation hierarchy in order to address early stages in the hierarchy (very small scales; high densities). Our second approach uses simulation-inspired results to describe the PDF that arises from lower-density larger-scale substructures which formed in more recent stages in the merger hierarchy. The distributions are skew positive, and they peak at densities lower than the mean density. The local dark-matter density may be as small as 1/10th the canonical value of $\\simeq0.4\\,\\GeV~\\cm^{-3}$, but it is probably no less than $0.2\\,\\GeV~\\cm^{-3}$. ", "introduction": "The nature of dark matter remains a mystery. Among the plethora of dark-matter candidates, there are two classes that are sufficiently promising to motivate major experimental searches. The first is a weakly interacting massive particles (WIMP) \\cite{Jungman:1995df,BHS05}, which may arise in supersymmetric extensions of the standard model or in theories that include universal extra dimensions (UED) \\cite{Hooper:2007qk}. The other leading candidate is the axion \\cite{axionreviews}, hypothesized to solve the strong-CP problem. WIMPs may be detected indirectly through observation of gamma rays or cosmic-ray positrons, antiprotons, and/or antideuterons produced when WIMPs annihilate in the Galactic halo or in the halos of some extragalactic systems. WIMPs might also be detected via observation of energetic neutrinos produced by annihilation of WIMPs that have accumulated in the Sun and/or Earth. There is also an effort to detect dark-matter particles in low-background experiments via detection of the $O(100\\,{\\rm keV})$ energy they impart to nuclei from which they elastically scatter. Likewise, dark-matter axions are being sought directly by conversion to photons in resonant-cavity experiments \\cite{axionexperiments}. If dark matter is composed of WIMPs, we may also get some clue to its nature from forthcoming LHC accelerator experiments \\cite{lhc}. Predictions for the event rates for any of these detection schemes depend on the distribution of dark matter in the Galactic halo. The flux of gamma and cosmic rays from WIMP annihilation depends on an integral of the square of the dark matter density over volume, while the rates for energetic neutrinos and direct detection depend only on the local dark-matter density. In early calculations, it was assumed simply that dark matter was smoothly distributed in the Galactic halo, with either an isothermal or NFW \\cite{NFW96} radial density profile, and with a local dark-matter density fixed by Milky Way dynamics to be $\\rhosun\\simeq0.4~\\GeV~\\cm^{-3}$. It has become clear in recent years, from theory and N-body simulations, that the distribution of dark matter in the Galactic halo is not likely to be perfectly smooth, and that some of the dark matter in the Galactic halo will be distributed into subhalos with a variety of sizes \\cite{substructure}. This substructure hierarchy may extend to very small scales \\cite{Boehm,Green:2003un,Diemand:2005vz,Diemand:2007qr,Ando:2005xg}. For WIMPs in supersymmetric and UED models, the cutoff scale of the power spectrum is in the range $[10^{-6} - 10^{2}\\, ]M_\\oplus$ \\cite{Chen:2001jz,Profumo:2006bv}. For axions, the growth of perturbations is suppressed on scales smaller than $(m_a H)^{-1/2}$ \\cite{Hu:2000ke,Boyle:2001du}, the geometric mean of the inverse of the axion mass and the Hubble constant; for example, for an axion of mass $m_a\\sim10^{-5}$~eV, the cutoff scale corresponds to about $10^{-11}\\,M_\\oplus$. The implications of such substructure for searches for cosmic rays from WIMP annihilation in the halo have been explored extensively in the literature \\cite{subgamma}. Simply stated, if a dark matter halo contains substructure, then the volume integral of the density squared is increased by some boost factor $B$. For example, Ref.~\\cite{Giocoli:2007gf} recently claimed $B\\sim10^3$ enhancements in the direction of the Galactic anticenter, although Ref.~\\cite{Berezinsky:2005py} finds an enhancement $B\\simeq2-5$. An analytic approach presented in Ref.~\\cite{Strigari:2006rd} also finds smaller values of $B$, of order $B \\sim 10$. The implications of substructure for {\\it direct} detection of dark matter have, however, been comparatively neglected. This is a serious omission, as substructure implies fluctuations in the local dark-matter density, and these fluctuations imply an uncertainty in the predicted rates for direct-detection experiments and for energetic-neutrino searches. A simple but illustrative example places all of the dark matter in subhalos of density $B$ times the mean density. In this case, the total annihilation rate is enhanced by a boost factor $B$, but the probability that the Solar System lives in such a subhalo is only $B^{-1}$. If $B\\gg 1$, this probability is small, implying dire prospects for direct-detection, an alarming conclusion that warrants further investigation. Few past studies focused on the implications of substructure for the local dark-matter density. Some focused on implications derived from numerical simulations \\cite{Moore:2001vq,Helmi:2002ss}, thus limited by resolution effects. Others looked at the contribution of tidal streams to the local density \\cite{Stiff:2001dq,Freese:2003tt}. The goal of this paper is to begin addressing questions such as: What are the possible values of the local density? How small can they be? What are the most probable values? To answer these questions, we calculate a probability distribution function (PDF) $P(\\rho)$ for the local dark-matter density. This PDF accounts for fluctuations in the local density due to substructure. The mean of this distribution is $\\simeq0.4~\\GeV~\\cm^{-3}$, the canonical value determined from dynamics, but we find that the local dark-matter density may be as small as 1/10th, but probably no less than half this canonical value. The PDF we calculate will have to be convolved with the results of null searches to infer constraints to particle-dark-matter parameters. In addition, the PDF can be used to assess what a substructure enhancement in the halo WIMP annihilation rate implies for the local dark matter density, and vice-versa. Our first approach, in Section II, uses an {\\it ansatz} about the survival fraction for subhalos that form early at high densities. General arguments will be used to bracket the plausible range of PDFs, and also the plausible lower limit to the local dark-matter density. This analytic approach capitalizes on the scale-invariant nature of the hierarchical formation of galaxies to extrapolate simulation results to the highest-density, smallest-scale substructures that form earliest and that are beyond the reach of simulations. We then perform a second calculation that assumes substructures with NFW profiles are distributed in the Milky Way halo. This second approach describes the local PDF due to larger-scale substructures which have formed from more recent stages in the merger hierarchy, and it thus complements the first approach. In Section III, we improve upon this latter calculation by using ingredients on subhalo mass functions and concentration parameters taken from numerical simulations. We compare our work to past literature in Section IV, and we state our conclusions in Section V. ", "conclusions": "It is evident from simulations and analytic arguments that some fraction of the local Milky Way dark matter may be in subhalos. The implications of substructure for indirect detection of WIMPs have been studied broadly, with the conclusion that there may be large enhancements in annihilation rates over the rates predicted assuming a smooth halo. The implications for {\\it direct} searches have, however, been largely overlooked. This is a possibly serious omission, as one consequence of substructure is that the local density will be smaller than the smoothly-distributed local density usually assumed. We have taken a few first steps to understand the implications of substructure for the local density. Our central goal is a calculation of the PDF $P(\\rho)$ for the local dark-matter density $\\rho$. This $P(\\rho)$ will need to be taken into account when interpreting the implications of null dark-matter searches for constraints to the particle-dark-matter parameter space (e.g., couplings and/or elastic-scattering cross sections). We considered two simple scenarios for substructure: In the first, early generations of very dense subhalos survive with some probability. The advantage of this approach is that if subhalo survival fractions can be measured in simulations for recently merged subhalos, the results might be extrapolated to the much earlier generations (much smaller and denser subhalos) that may be below the resolution of simulations. In the second approach, the halo is assumed to consist of recently formed subhalos, each with an NFW profile. This approach provides a simple, albeit approximate, way to understand the effects of larger-scale substructure, from more recent stages in the merger hierarchy, on the PDF. We then pursued this approach further using subhalo mass functions and concentration-parameter distributions taken from simulations. This approach does not take into account the possible contribution of subhalos within subhalos. In principle, a complete solution for the PDF, including substructures on the largest and smallest scales, can be obtained by convolving our two calculations. This, however, will be left for future research. Substructure scenarios that yield larger annihilation enhancements generally imply a smaller local dark-matter density. Very large annihilation enhancements require that the very densest substructures, which generally form earlier, must survive through all later generations of structure formation. If earlier substructures are less likely to survive than more recent substructures, then a very large annihilation enhancement is unlikely. So, how small can the local density be? The smallest local density in the models we surveyed was one tenth the canonical value of $0.4\\, \\GeV\\,\\cm^{-3}$, usually assumed for a smoothly distributed halo. This small value was obtained from our simulation-inspired result using what we believe to be an overly conservative estimate of the smooth fraction, which we obtained by truncating NFW subhalos when their density falls below the mean halo density. More realistic values for the smooth fraction are probably in the range of 50\\%-80\\%, which would then correspond in our simulation-inspired results to the $\\xi=0.8$ and $\\xi=0.5$ distributions, respectively, shown in Fig.~\\ref{fig:fig1}. The smallest local density in our analytic model for early substructure was 0.3, obtained using the high value, $f(\\rhosun)=0.2$, and assuming a constant $f(\\rho_1)$. If, however, $f(\\rhosun)$ is lower, then the local density is increased. More importantly, it is quite likely that $f(\\rho)$ decreases with $\\rho$, and if so, the local density is increased to $\\sim0.8$ times its canonical value, even for $f(\\rhosun)=0.2$, and even higher for smaller $f(\\rhosun)$. A combination of our two approaches, to take into account both low- and high-density substructures, will likely show that the local density is no less than half the canonical value. The volume of the halo probed during a three-year direct-detection experiment is very small, and so the halo density $\\rho$ that we have been discussing can be safely assumed to be the density averaged over the duration of a direct-detection experiment. However, the rate for production of energetic neutrinos from WIMP annihilation in the Sun or Earth lags behind the rate for capture of WIMPs from the halo by an equilibration time $t_{\\rm eq}$ \\cite{Griest:1986yu,Kamionkowski:1991nj} that can vary considerably with the WIMP's mass and elastic-scattering and annihilation cross sections; typical values might be $t_{\\rm eq}^{\\odot}\\sim 5\\times 10^7$ yr and $t_{\\rm eq}^\\oplus\\sim10^{10}$ yr \\cite{Kamionkowski:1994dp}. An energetic-neutrino search thus probes the halo density averaged over a much larger volume (for the latest bounds on the flux of high-energy neutrinos from annihilations in the Earth and the Sun see \\cite{Desai:2004pq,Ackermann:2005fr, Achterberg:2006jf}).The halo density averaged over this volume will thus have a PDF that will be narrower than that for direct detection, and the minimum density will be closer to the mean halo density. However, even though initially it was assumed that the velocity distribution function of dark matter particles mirrors that in free space (thus justifying the use of a Maxwell-Boltzmann distribution) \\cite{Gould:1991}, it is possible that the velocity distribution function is surpressed in the low-velocity tail due to the effects of solar capture and WIMP diffusion in the solar system due to the presence of other planets \\cite{Gould:1999je,Lundberg:2004dn}. This surpression manifests itself as a reduction in the number density of dark matter particles near the Earth for WIMPs with masses greater than few hundred GeV. Nevertheless, there is a possibility that for low-mass WIMPs, given the longer equilibration time for the Earth relative to that for the Sun, the annihilation signal from the Earth could be boosted relative to that for the Sun, if the Solar System passed through a very dense subhalo at a time $t_{\\rm eq}^\\odot 0.35$), while evolution of systems with more evolved donors with masses $\\simeq 1\\,\\ms$ opens possibility to explain the origin of LMBHB with longer \\per. 2. An alternative model suggests that if accretion discs are truncated in quiescence they could be cold and stable. Then truncation of discs at radii close to $R_{\\rm circ}$ can make discs in virtually all LMBHB cold and stable. In this case SXT outbursts perhaps are not related to DIM and are random events that result from \"external factors\", like episodes of enhanced mass transfer or variability in truncation radius. 3. Population synthesis models presented above are obtained under assumption of a high value of the product of binding energy parameter of stellar envelopes and expulsion efficiency of common envelopes -- \\al=2. Model with \\al=0.5 gives similar results (but a reduced total number of systems). Further decrease of \\al\\ results in models that are not consistent with observed SXT. The issue of \\al\\ is still controversial \\citep[see, e.g. ][]{Taudew,Podsiadlowski&al03} in the sense that there are no strict criteria for defining binding energy of stellar envelopes and there is no clear understanding whether sources other than gravitational energy may contribute to unbinding common envelopes. Apart of common envelopes, crucial role in determining the possibility of formation of LMBHB is played by stellar wind mass-loss in all stages of stellar evolution, since it defines parameters of the system at the beginning of common envelope stage and immediately before supernova explosion that gives birth to black hole. If general notions on mass-loss will not be revised, our results suggest that existence of short-period low-mass X-ray binaries demands high values of \\al\\ parameter." }, "0801/0801.3427_arXiv.txt": { "abstract": "Several $\\gamma$-ray binaries have been recently detected by the High-Energy Stereoscopy Array (H.E.S.S.) and the Major Atmospheric Imaging Cerenkov (MAGIC) telescope. In at least two cases, their nature is unknown. In this paper we aim to provide the details of a theoretical model of close $\\gamma$-ray binaries containing a young energetic pulsar as compact object, earlier presented in recent Letters. This model includes a detailed account of the system geometry, the angular dependence of processes such as Klein-Nishina inverse Compton and $\\gamma\\gamma$ absorption in the anisotropic radiation field of the massive star, and a Monte Carlo simulation of leptonic cascading. We present and derive the used formulae and give all details about their numerical implementation, particularly, on the computation of cascades. In this model, emphasis is put in the processes occurring in the pulsar wind zone of the binary, since, as we show, opacities in this region can be already important for close systems. We provide a detailed study on all relevant opacities and geometrical dependencies along the orbit of binaries, exemplifying with the case of LS 5039. This is used to understand the formation of the very high-energy lightcurve and phase dependent spectrum. For the particular case of LS 5039, we uncover an interesting behavior of the magnitude representing the shock position in the direction to the observer along the orbit, and analyze its impact in the predictions. We show that in the case of LS 5039, the H.E.S.S. phenomenology is matched by the presented model, and explore the reasons why this happens while discussing future ways of testing the model. ", "introduction": "Very recently, a few massive binaries have been identified as variable very-high-energy (VHE) $\\gamma$-ray sources. They are PSR B1259-63 (Aharonian et al. 2005a), LS 5039 (Aharonian et al. 2005b, 2006), LS I +61 303 (Albert et al. 2006, 2008a,b), and Cyg X-1 (Albert et al. 2007). The nature of only two of these binaries is considered known: PSR B 1259-63 is formed with a pulsar whereas Cyg X-1 is formed with a black hole compact object. The nature of the two remaining systems is under discussion. The high-energy phenomenology of Cyg X-1 is different from that of the others. It has been detected just once in a flare state for which a duty cycle is yet unknown. The three other sources, instead, present a behavior that is fully correlated with the orbital period. The latter varies from about 4 days in the case of LS 5039 to several years in the case of PSR B1259-63: this span of orbital periodicities introduces its own complications in analyzing the similarities among the three systems. LS I +61 303 shares with LS 5039 the quality of being the only two known microquasars/$\\gamma$-ray binaries that are spatially coincident with sources above 100 MeV listed in the Third Energetic Gamma-Ray Experiment (EGRET) catalog (Hartman et al. 1999). These sources both show low X-ray emission and variability, and no signs of emission lines or disk accretion. For LS I +61 303, extended, apparently precessing, radio emitting structures at angular extensions of 0.01-0.05 arcsec have been reported by Massi et al. (2001, 2004); this discovery has earlier supported its microquasar interpretation. But the uncertainty as to what kind of compact object, a black hole or a neutron star, is part of the system (e.g., Casares et al. 2005a), seems settled for many after the results presented by Dhawan et al. (2006). These authors have presented observations from a July 2006 VLBI campaign in which rapid changes are seen in the orientation of what seems to be a cometary tail at periastron. This tail is consistent with it being the result of a pulsar wind. Indeed, no large features or high-velocity flows were noted on any of the observing days, which implies at least its non-permanent nature. The changes within 3 hours were found to be insignificant, so the velocity can not be much over 0.05$c$. Still, discussion is on-going (e.g. see Romero et al. 2007, Zdziarski et al. 2008). New campaigns with similar radio resolution, as well as new observations in the $\\gamma$-ray domain have been obtained since the Dhawan's et al. original results (Albert et al. 2008b). A key aspect in these high-angular-resolution campaigns is the observed maintenance in time of the morphology of the radio emission of the system: the changing morphology of the radio emission along the orbit would require a highly unstable jet, which details are not expected to be reproduced orbit after orbit as indicated by current results (Albert et al. 2008b). The absence of accretion signatures in X-rays in Chandra and XMM-Newton observations (as reported by Sidoli et al. 2006, Chernyakova et al. 2006, and Paredes et al. 2007) is another relevant aspect of the discussion about the compact object companion. Finally, it is interesting to note that neutrino detection or non-detection with ICECUBE will shed light on the nature of the $\\gamma$-ray emission irrespective of the system composition (e.g., Aharonian et al. 2006b, Torres and Halzen 2007). For LS 5039, a periodicity in the $\\gamma$-ray flux, consistent with the orbital timescale as determined by Casares et al. (2005b), was found with amazing precision (Aharonian et al. 2006). Short timescale variability displayed on top of this periodic behavior, both in flux and spectrum, was also reported. It was found that the parameters of power-law fits to the $\\gamma$-ray data obtained in 0.1 phase binning already displayed significant variability. Current H.E.S.S. observations of LS 5039 ($\\sim 70$ hours distributed over many orbital cycles, Aharonian et al. 2006) constitute one of the most detailed datasets of high-energy astrophysics. Similarly to LS I +61 303, the discovery of a jet-like radio structure in LS 5039 and the fact of it being the only radio/X-ray source co-localized with a mildly variable (Torres et al. 2001a,b) EGRET detection, prompted a microquasar interpretation (advanced already by Paredes et al. 2000). However, the current mentioned findings at radio and VHE $\\gamma$-rays in the cases of LS I +61 303 (Dhawan et al. 2006, Albert et al. 2006, 2008b) or PSR B1259-63 (Aharonian et al. 2005), gave the perspective that all three systems are different realizations of the same scenario: a pulsar-massive star binary. Dubus (2006a,b) has studied these similarities. He provided simulations of the extended radio emission of LS 5039 showing that the features found in high resolution radio observations could also be interpreted as the result of a pulsar wind. Recently, Rib\\'o et al. (2008) provided VLBA radio observations of LS 5039 with morphological and astrometric information at milliarcsecond scales. They showed that a microquasar scenario cannot easily explain the observed changes in morphology. All these results, together with the assessment of the low X-ray state (Martocchia et al. 2005) made the pulsar hypothesis tenable, and the possibility of explaining the H.E.S.S. phenomenology in such a case, an interesting working hypothesis. High energy emission from pulsar binaries has been subject of study for a long time (just to quote a non-exhaustive list of references note the works of Maraschi and Treves 1981; Protheroe and Stanev 1987, Arons and Tavani 1993, 1994; Moskalenko et al. 1993; Bednarek 1997, Kirk et al. 1999, Ball and Kirk 2000, Romero et al. 2001, Anchordoqui et al. 2003, and others already cited above). LS 5039 has been recently subject of intense theoretical studies (e.g., Bednarek 2006, 2007; which we comment on in more detail below, Bosch-Ramon et al. 2005; B\\\"ottcher 2007; B\\\"ottcher and Dermer 2005; Dermer and B\\\"ottcher 2006; Dubus 2006a,b; Paredes et al. 2006; Khangulyan et al. 2007; Dubus et al. 2007). In the penultimate paper mentioned in the list above, Khangulyan et al. (2007), and contrary to the assumption here, authors assumed a jet structure perpendicular to the orbital plane of the system. The energy spectrum and lightcurves were computed, accounting for the acceleration efficiency, the location of the accelerator along the jet, the speed of the emitting flow, the inclination angle of the system, as well as specific features related to anisotropic inverse Compton (IC) scattering and pair production. Different magnetic fields, affecting Synchrotron emission, and the losses they produced, were also tested given a large model parameter space. Authors found a good agreement between H.E.S.S. data for some of their models. In the last of these papers, Dubus et al. (2007) computed the phase dependent lightcurve and spectra expected from inverse Compton interactions from electrons injected close to the compact object, assumed as a likely rotation-powered pulsar. Since the angle at which an observer sees the binary and propagating electrons changes with the orbit (see below), a phase dependence of the spectrum is expected, and anisotropic inverse Compton is needed to compute it. In general, they found that the lightcurve is a good fit to the observations, except at the phases of maximum attenuation where pair cascade emission plays a role. Dubus et al. (2007) do not consider cascading in their models, as we do here. Without cascading, zero flux is expected at a broad phase around periastron, which is not found. This lack of cascading in their model also affects the spectra, which are not reproduced well, particularly at the superior conjunction broad phases of the orbit. They mentioned that both, cascading and/or a change in the slope of the power-law injection for the interacting electron distribution could be needed to explain the spectrum in these phases, what we explore in detail in this work. In order to compute inverse Compton emission from LS 5039, we use, as in previous works, leptons interacting with the star photon field. Geometry is described there with different levels of detail, what influence the results. In general, cascading processes were not taken into account, and the goodness of fitting the H.E.S.S. data is arguable in most cases, both for the lightcurve and spectrum. In none of the papers mentioned above, the theoretical predictions for the short timescale spectral variability found by H.E.S.S. in 0.1 phase binning was shown and compared with data. We discuss these results from our model below. In recent Letters (Sierpowska-Bartosik and Torres 2007, 2008) under the assumption that LS 5039 is composed by a pulsar rotating around an O6.5V star in the $\\sim 3.9$ day orbit, we presented the results of a leptonic (for a generic hadronic model see Romero et al. 2003) theoretical modeling for the high-energy phenomenology observed by H.E.S.S. These works studied the lightcurve, the spectral orbital variability in both broad orbital phases and in shorter (0.1 phase binning) timescales and have found a complete agreement between H.E.S.S. observations and our predictions. We have also analyzed how this model could be tested by Gamma-ray Large Area Space Telescope (GLAST), and how much time would be needed for this satellite in order to rule the model out in case theory significantly departs from reality. But many details of implementation which are not only useful for the case of LS 5039 but for all others close massive $\\gamma$-ray binaries, as well as many interesting results concerning the binary geometry, wind termination, opacities to different processes along the orbit of the system, and further testing at the highest energy $\\gamma$-ray domain were left without discussion in our previous works. Here, we provide these details, together with benchmark cases that are useful to understand the formation of the very high-energy lightcurve and phase dependent spectra. The rest of this paper is organized as follows. Next Section introduces the model concept and its main properties. It provides a discussion of geometry, wind termination, and opacities along the orbit of the system (we focus on LS 5039). An accompanying Appendix provides mathematical derivations of the formulae used and useful intermediate results that are key for the model, but too cumbersome to include them as part of the main text. It also deals with numerical implementation, and describes in detail the Monte Carlo simulation of the cascading processes. The results follow: Section 3 deals with a mono-energetic interacting particle population, and Section 4, with power-law primary distributions. Comparison with H.E.S.S. results is made in these Sections and details about additional tests are given. Final concluding remarks are provided at the end. ", "conclusions": "We presented the details of a theoretical model for the high-energy emission from close $\\gamma$-ray binaries, and applied it to the particular case of LS 5039. The model assumes a pulsar scenario, where either the pulsar or a close-to-the-pulsar shock injects leptons that after being reprocessed by losses to constitute a steady population, are assumed to interact with the target photon field provided by the companion star within the PWZ. The model accounts for the highly variable system geometry with respect to the observer, and radiative processes; essentially, anisotropic Klein-Nishina ICS and $\\gamma\\gamma$ absorption, put together with a Monte Carlo computation of cascading. The formation of lightcurve and spectra in this model was discussed in detail for the case where the interacting leptons are assumed mono-energetic and described by power-laws. Comparing the interacting models of mono-energetic leptons and power-law distributions we can see similar dependencies for the spectral changes along the orbit and the GeV to TeV lightcurves. Thus, the case of a mono-energetic population is useful to understand some of the aspects regarding the formation of the observational features, although it does not match observational data. For the mono-energetic case, we have shown the spectra produced at characteristic phases (INFC, SUPC, periastron, apastron) for primary energies of 1 TeV and 10 TeV. In the power-law distribution model, the specific features of the photon spectra and the lightcurves produced for an specific primary electron energy overlap. We can still notice the absorption features in the spectra produced at SUPC. We have discussed effects solely based on the optical depths and the general geometrical dependence along the orbit, as well as on the presence of the shock which terminates the pulsar wind in the direction to the observer at SUPC phases. A power-law lepton distribution interacting in the PWZ describes very well the phenomenology found in the LS 5039 system at all timescales, both flux and spectrum-wise, even at the shortest timescales measured. This latter result is unexpected: we find that there is nothing a priori in the model that allows one to predict that when broad phase spectra data (INFC and SUPC) are reproduced so will be the data at the individual and much shorter phase-binning, less with such a good agreement. This result point perhaps to some reliability of the model, at least in its essential ingredients: geometry, cascading, interacting electron population. However, this model certainly has room for improvement. We emphasize here that we do not have an a priori model for the interacting lepton population itself, although we have discussed the research on dissipation processes in the PWZ which may give raise to such distributions if it results from pulsar injection. In any case, the assumption of power-laws is an approximation to a more complex scenario where the real interacting lepton population is the result of a full escape-loss equation. In addition, we are not considering yet the multiwavelength emission at lower energies, since we left out of our description the synchrotron emission of electrons accelerated at the shock and the morphology of the shock along the orbit, what we expect to discuss elsewhere. Other than system scalings that are fixed by multiwavelength observations, the model is based on just a handful of free parameters, and it is subject to tests at high and very high-energy $\\gamma$-ray observations with both GLAST (described in more detail in Sierpowska-Bartosik \\& Torres 2008) and future samples of data at higher energies, where more statistics at finer phase bins can determine better the spectral evolution along the orbit of this interesting system.\\\\ {\\sl We acknowledge extended use of IEEC-CSIC parallel computers cluster. We acknowledge W. Bednarek for discussions, and the Referee for useful comments. This work was supported by grants AYA 2006-00530 and CSIC-PIE 200750I029.} \\newpage" }, "0801/0801.4031_arXiv.txt": { "abstract": "We revisit the distant future of the Sun and the solar system, based on stellar models computed with a thoroughly tested evolution code. For the solar giant stages, mass-loss by the cool (but not dust-driven) wind is considered in detail. Using the new and well-calibrated mass-loss formula of Schr\\\"oder \\& Cuntz (2005, 2007), we find that the mass lost by the Sun as an RGB giant (0.332 $M_{\\odot}$, 7.59 Gy from now) potentially gives planet Earth a significant orbital expansion, inversely proportional to the remaining solar mass. According to these solar evolution models, the closest encounter of planet Earth with the solar cool giant photosphere will occur during the tip-RGB phase. During this critical episode, for each time-step of the evolution model, we consider the loss of orbital angular momentum suffered by planet Earth from tidal interaction with the giant Sun, as well as dynamical drag in the lower chromosphere. As a result of this, we find that planet Earth will not be able to escape engulfment, despite the positive effect of solar mass-loss. In order to survive the solar tip-RGB phase, any hypothetical planet would require a present-day minimum orbital radius of about 1.15 AU. The latter result may help to estimate the chances of finding planets around White Dwarfs. Furthermore, our solar evolution models with detailed mass-loss description predict that the resulting tip-AGB giant will not reach its tip-RGB size. Compared to other solar evolution models, the main reason is the more significant amount of mass lost already in the RGB phase of the Sun. Hence, the tip-AGB luminosity will come short of driving a final, dust-driven superwind, and there will be no regular solar planetary nebula (PN). The tip-AGB is marked by a last thermal pulse and the final mass loss of the giant may produce a circumstellar (CS) shell similar to, but rather smaller than, that of the peculiar PN IC 2149 with an estimated total CS shell mass of just a few hundredths of a solar mass. ", "introduction": "Climate change and global warming may have drastic effects on the human race in the near future, over human time-scales of decades or centuries. However, it is also of interest, and of relevance to the far future of all living species, to consider the much longer-term effects of the gradual heating of the Earth by a more luminous Sun as it evolves towards its final stage as a white dwarf star. This topic has been explored on several occasions (e.g. Sackmann, Boothroyd \\&\\ Kraemer 1993, Rybicki \\&\\ Denis 2001, Schr\\\"oder, Smith \\&\\ Apps 2001 (hereafter SSA)), and has been discussed very recently by Laughlin (2007). Theoretical models of solar evolution tell us that the Sun started on the zero-age main sequence (ZAMS) with a luminosity only about 70\\%\\ of its current value, and it has been a long-standing puzzle that the Earth seems none the less to have maintained a roughly constant temperature over its life-time, in contrast to what an atmosphere-free model of irradiation would predict. Part of the explanation may be that the early atmosphere, rich in CO$_2$ that was subsequently locked up in carbonates, kept the temperature up by a greenhouse effect which decreased in effectiveness at just the right rate to compensate for the increasing solar flux. The r\\^ole of clouds, and their interaction with galactic cosmic rays (CR), may also be important: there is now some evidence (Svensmark 2007; but see Harrison et al. 2007 and Priest et al. 2007) that cosmic rays encourage cloud cover at low altitudes, so that a higher CR flux would lead to a higher albedo and lower surface temperature. The stronger solar wind from the young Sun would have excluded galactic cosmic rays, so cloud cover on the early Earth may have been less than now, allowing the full effect of the solar flux to be felt. What of the future? Although the Earth's atmosphere may not be able to respond adequately on a short time-scale to the increased greenhouse effect of carbon dioxide and methane released into the atmosphere by human activity, there is still the possibility, represented by James Lovelock's Gaia hypothesis (Lovelock 1979, 1988, 2006), that the biosphere may on a longer time-scale be able to adjust itself to maintain life. Some doubt has been cast on that view by recent calculations (Scaife, private communication, 2007; for details, see e.g. Cox et al. 2004, Betts et al. 2004) which suggest that, on the century timescale, the inclusion of biospheric processes in climate models actually leads to an increase in carbon dioxide emissions, partly through a feedback that starts to dominate as vegetation dies back. In any case, it is clear that the time will come when the increasing solar flux will raise the mean temperature of the Earth to a level that not even biological or other feedback mechanisms can prevent. There will certainly be a point at which life is no longer sustainable, and we shall discuss this further in Section \\ref{habitablezone}. After that, the fate of the Earth is of interest mainly insofar as it tells us what we might expect to see in systems that we observe now at a more advanced stage of evolution. We expect the Sun to end up as a white dwarf -- do we expect there to be any planets around it, and in particular do we expect any small rocky planets like the Earth? The question of whether the Earth survives has proved somewhat tricky to determine, with some authors arguing that the Earth survives (e.g. SSA) and others (e.g. Sackmann et al. 1993) claiming that even Venus survives, while general textbooks (e.g. Prialnik 2000, p.10) tend to say that the Earth is engulfed. A simple model (e.g. SSA), ignoring mass loss from the Sun, shows clearly that all the planets out to and including Mars are engulfed, either at the red giant branch (RGB) phase -- Mercury and Venus -- or at the later asymptotic giant branch (AGB) phase -- the Earth and Mars. However, the Sun loses a significant amount of mass during its giant branch evolution, and that has the effect that the planetary orbits expand, and some of them keep ahead of the advancing solar photosphere. The effect is enhanced by the fact (SSA) that when mass loss is included the solar radius at the tip of the AGB is comparable to that at the tip of the RGB, instead of being much larger; Mars certainly survives, and it appears (SSA) that the Earth does also. The crucial question here is: what is the rate of mass loss in real stars? Ultimately this must be determined from observations, but in practice these must be represented by some empirical formula. Most people use the classical Reimers' formula (Reimers 1975, 1977), but there is considerable uncertainty in the value to be used for his parameter $\\eta$, and different values are needed to reproduce the observations in different parameter regimes. In our own calculations (SSA) we used a modification of the Reimers' formula, which has since been further improved and calibrated rather carefully against observation, so that we believe that it is currently the best available representation of mass loss from stars with non-dusty winds (Schr\\\"oder \\&\\ Cuntz 2005, 2007 -- see Section \\ref{evolution}, where we explore the consequences of this improved mass-loss formulation). However, although we have considerably reduced the uncertainties in the mass-loss rate, there is another factor that works against the favourable effects of mass loss: tidal interactions. Expansion of the Sun will cause it to slow its rotation, and even simple conservation of angular momentum predicts that by the time the radius has reached some 250 times its present value (cf. Table \\ref{solarmodels}) the rotation period of the Sun will have increased to several thousand years instead of its present value of under a month; effects of magnetic braking will lengthen this period even more. This is so much longer than the orbital period of the Earth, even in its expanded orbit, that the tidal bulge raised on the Sun's surface by the Earth will pull the Earth back in its orbit, causing it to spiral inwards. This effect was considered by Rybicki \\&\\ Denis (2001), who argued that Venus was probably engulfed, but that the Earth might survive. An earlier paper by Rasio et al. (1996) also considered tidal effects and concluded on the contrary that the Earth would probably be engulfed. However, the Rybicki \\&\\ Denis calculations were based on combining analytic representations of evolution models (of Hurley, Pols \\&\\ Tout 2000) with the original Reimers' mass-loss formula rather than on full solar evolution calculations with a well-calibrated mass-loss formulation. The Rasio et al. paper also employed the original Reimers' formula, and both papers use somewhat different treatments of tidal drag. We have therefore re-considered this problem in detail, with our own evolutionary calculations and an improved mass-loss description as the basis; full details are given in Sections \\ref{evolution} and \\ref{tiprgb}. ", "conclusions": "We have applied an improved and well-tested mass-loss relation to RGB and AGB solar evolution models, using a well-tested evolution code. While the habitable zone in the inner solar system will already move outwards considerably in the next 5 billion years of solar MS evolution, marking the end of life on Earth, the most critical and fatal phase for the inner planetary system is bound to come with the final ascent of the Sun to the tip of the RGB. Considering in detail the loss of angular momentum by tidal interaction and dynamical drag in the lower chromosphere of the solar giant, we have been able to compare the evolution of the RGB solar radius with that of the orbit of planet Earth. Our computations reveal that planet Earth will be engulfed by the tip-RGB Sun, just half a million years before the Sun will have reached its largest radius of 1.2 AU, and 1.0 (3.8) million years after Venus (and Mercury) have suffered the same fate. While solar mass loss alone would allow the orbital radius of planet Earth to grow sufficiently to avoid this ``doomsday'' scenario, it is mainly tidal interaction of the giant convective envelope with the closely orbiting planet which will lead to a fatal decrease of its orbital size. The loss of about 1/3 of the solar mass already on the RGB has significant consequences for the solar AGB evolution. The tip-AGB Sun will not qualify for an intense, dust-driven wind and, hence, will not produce a regular PN. Instead, an insubstantial circumstellar shell of just under 1/100\\,$M_{\\odot}$ will result, and perhaps a peculiar PN similar to IC\\,2149." }, "0801/0801.1422_arXiv.txt": { "abstract": "We present a spatially resolved 894\\,$\\mu$m map of the circumstellar disk of the Butterfly star in Taurus (IRAS\\,04302+2247), obtained with the Submillimeter Array (SMA). The predicted and observed radial brightness profile agree well in the outer disk region, but differ in the inner region with an outer radius of \\about80-120\\,AU. In particular, we find a local minimum of the radial brightness distribution at the center, which can be explained by an increasing density / optical depth combined with the decreasing vertical extent of the disk towards the center. Our finding indicates that young circumstellar disks can be optically thick at wavelengths as long as 894\\,$\\mu$m. While earlier modeling lead to general conclusions about the global disk structure and, most importantly, evidence for grain growth in the disk (Wolf, Padgett, \\& Stapelfeldt~2003), the presented SMA observations provide more detailed constraints for the disk structure and dust grain properties in the inner, potentially planet-forming region (\\aboutless80-120\\,AU) vs.\\ the outer disk region (\\about120-300\\,AU). ", "introduction": "IRAS~04302+2247 is a Class~I protostar in the Taurus-Auriga molecular cloud complex whose equatorial plane is inclined edge-on to the line of sight (inclination = 90$^{\\rm{o}} \\pm 3^{\\rm{o}}$; Wolf et al.~2003, hereafter WPS03). Parallel to the increasing amount of observational constraints, such as ground-based near-infrared images and polarization maps (Lucas \\& Roche~1997, 1998), near-infrared images obtained with the Hubble Space Telescope (Padgett et al.~1999) and spatially resolved 1.3\\,mm and 2.7\\,mm maps (WPS03), several attempts have been undertaken to model the structure and physical conditions in the circumstellar disk and envelope of this object (e.g., Lucas \\& Roche~1998, WPS03, Stark et al.~2006). In particular, spatially resolved images of the circumstellar environment of the Butterfly Star, obtained in the near-infrared and millimeter wavelength range, allowed WPS03 to conclude that the grains in the envelope of this object cannot be distinguished from those of the interstellar medium, while grains have grown via coagulation by up to 2 - 3 orders of magnitude in the much denser circumstellar disk. The separated dust grain evolution is in agreement with the theoretical prediction of a sensitive dependence of grain growth on the location in the circumstellar environment of young (proto)stars: Grain growth is expected to occur on much shorter timescales in the dense region of circumstellar disks than in the thin circumstellar envelope. For the same reason a radial dependence of the dust grain evolution in the disk itself is expected. However, the observational data presented by WPS03 did not allow to constrain the spatial dependence of the dust grain properties in the disk. Based on radiative transfer simulations, using the disk model by WPS03, we found that further insights into and constraints for the dust grain growth as the first stage of planet formation in the circumstellar disk of the Butterfly Star can be obtained with high-resolution submillimeter observations. As outlined in Sect.~\\ref{sect.ana}, the apparent structure of the disk is predicted to change significantly as the observing wavelength is decreased from millimeter to submillimeter wavelengths, allowing to constrain the radial and vertical disk structure and distribution of the dust grain properties. In this paper we present and discuss new, spatially resolved observations of the circumstellar disk of the Butterfly Star, obtained with the Submillimeter Array\\footnote{ The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory and the Academia Sinica Institute of Astronomy and Astrophysics and is funded by the Smithsonian Institution and the Academia Sinica.} (SMA) at 894\\,$\\mu$m. The observations and data reduction are described in Sect.~\\ref{sect.obs}, followed by a description of the data analysis (Sect.~\\ref{sect.ana}) and conclusions in Sect.~\\ref{sect.sum}. ", "conclusions": "\\label{sect.sum} We obtained the first spatially resolved submillimeter map of circumstellar disk of the prominent Butterfly Star in Taurus. We find a good agreement between the observed and the predicted brightness profile in the outer region of the disk. However, a discrepancy between the predicted and observed radial flux distribution was found in the inner region (inside \\about80-120\\,AU), showing a decrease of the flux towards the center. This discrepancy -- a local minimum at the stellar position -- amounts to only \\about2$\\sigma$. However, the specific location of the minimum at the disk center and the slight symmetry of the corresponding local maxima with respect to the disk center indicate that this profile is an intrinsic feature of the real disk brightness distribution. Based on radiative transfer simulations and additional observations at 1.36\\,mm we exclude a large inner hole (strongly depleted from small dust grains) as a possible explanation of this local minimum in the radial brightness distribution. These observations may provide the basis for a detailed model of the inner structure of the disk of the Butterfly Star. While earlier modeling lead to conclusion about grain growth in the disk (WPS03), the new SMA observations provide constraints for the disk structure and dust grain properties in the inner, potentially planet-forming region (\\about80-120\\,AU) vs.\\ the outer disk region (\\about120-300\\,AU). It was found that the optical depth in the inner disk region is higher than predicted by WPS03. The column density of dust along the line of sight is therefore higher in the inner disk region than derived from the previous disk model in which a perfect mixing of dust and gas throughout the entire disk is assumed. These observations indicate that there exists a higher dust density and therefore a higher dust-to-gas mass ratio in the inner disk region. This conclusion is based on the assumption that the density profile of the gas phase of the disk can be described by the same approach in the inner and outer disk region (Eq.~\\ref{dendisk}) and that there exists no discontinuity in the gas density profile. Since the optical depth effect discussed above is constrained by both the radial and vertical disk structure and dust grain properties, such a model will have to take the predicted grain evolution and its dependence on the radial and vertical position in the disk into account. Furthermore, dust settling and the resulting increase of the grain-grain interaction probability will result in a further vertical dependence of the grain size distribution (e.g., Weidenschilling~1997). Beside the grain evolution, mixing processes, such as convection and radial mixing within circumstellar disks have to be considered since these processes lead to a redistribution of processed, evolved dust grains to outer, less dense and colder disk regions (see, e.g., Klahr et al.~1999, Gail~2001). Our observations illustrate the high potential of submillimeter observations for studying circumstellar disks around young (proto)stars. Although the spatial resolution is significantly lower than that aimed for in the case of ALMA in a few years from now, the tight correlation between the density and temperature structure in the disk is already able to constrain the radial and vertical disk structure and thus the dust grain properties as a function of the distance from the disk midplane, based on the spatially resolved radial brightness profile obtained with the SMA in the submillimeter wavelength range." }, "0801/0801.1108_arXiv.txt": { "abstract": "Passive spiral galaxies, despite their spiral morphological appearance, do not have any emission lines indicative of ongoing star formation in their optical spectra. Previous studies have suggested that passive spiral galaxies preferentially exist in infall regions of galaxy clusters, suggesting that the cluster environment is likely to be responsible for creating these galaxies. By carrying out spatially resolved long-slit spectroscopy on four nearby passive spiral galaxies with the Apache Point Observatory 3.5-m telescope, we investigated the stellar populations of passive spiral galaxies separately for their inner and outer regions. In the two unambiguously passive spiral galaxies among the four observed galaxies, H$\\delta$ absorption lines are more prominent in the outer regions of the galaxies, whereas the 4000-{\\AA} breaks (D$_{4000}$) are strongest in the inner regions of the galaxies. A comparison with a simple stellar population model for the two passive spiral galaxies indicates that the outer regions of the galaxies tend to harbour younger populations of stars. The strong H$\\delta$ absorption observed in the outer regions of the sample galaxies is consistent with that of galaxies whose star formation ceased a few Gyrs ago. Because of the large uncertainty in the absorption indices in our samples, further observations are needed in order to place constraints on the mechanisms that quench star formation in passive spiral galaxies. ", "introduction": "\\label{sec:section1} Passive spiral galaxies have peculiar spectroscopic characteristics among the galaxy populations having spiral morphologies. They show few or neither of the emission lines in H$\\alpha$ nor [O II] in optical spectra that would be indicative of ongoing star formation (Couch et al. 1998; Dressler et al. 1999; Poggianti et al. 1999; Goto et al. 2003b). The optical g $-$ r colours observed in such galaxies are found to be significantly redder than those of spiral galaxies with emission lines, which confirms the lower star formation rate among passive spiral galaxies (Poggianti et al. 1999). A similar population of galaxies, known as 'anemic' spiral galaxies, has been found in the Virgo Cluster (van den Bergh 1976). They have smoothed spiral arms showing less prominent star formation activity than in other galaxies of the same Hubble type. Elmegreen et al. (2002) observed anemic spirals in the Virgo Cluster and found lower gas surface densities than for normal spirals for these galaxies. They showed that the gas surface density in these galaxies is below the threshold for star formation (Kennicutt 1989), suggesting that the lack of star formation is caused by the stripping of gas in the environment of the cluster (Elmegreen et al. 2002). In order to investigate the effect of environment on the formation of such galaxies, Goto et al. (2003b) searched for passive spiral galaxies in all environments, including in dense cluster cores and field regions, using the large samples in the Sloan Digital Sky Survey (SDSS) data (Strauss et al. 2002). They found that passive spiral galaxies preferentially exist in the environment where a local galaxy density is intermediate between that of the cluster cores and the field regions (Goto et al. 2003b). This characteristic environment corresponds to the region where a significant decline in the star formation rate has been observed (Lewis et al. 2002; Gomez et al. 2003; Tanaka et al. 2004). These studies imply that cluster-related phenomena could be the main factors responsible for the formation of passive spiral galaxies. Recent observations have revealed that the cluster environment may bring about a transformation in the galaxy population from star-forming spirals to passive galaxies, and thus passive spiral galaxies have been suggested to be a population in a transitional phase between these two populations (Poggianti et al. 1999). For instance, the fraction of blue galaxies is found to be larger in distant clusters than in local clusters, a phenomenon known as the Butcher-Oemler effect (Butcher \\& Oemler 1978; Couch et al. 1998; Kodama \\& Bower 2001; Goto et al. 2003d, 2004). The Butcher-Oemler effect implies that the environment of a cluster may affect the star formation activity of its member galaxies. Furthermore, morphological type (Goto et al. 2003c; Treu et al. 2003; Postman et al. 2005) is found to be correlated with various functions of cluster environment, such as local density or cluster centric radius (Goto et al. 2004; Tanaka et al. 2004). Numerical simulations (Bekki et al. 2002) have shown that gas-stripped galaxies may finally become S0 galaxies if no further accretion onto the disc occurs after the stripping. UV observations suggest that the cessation of star formation could take place before this morphological transformation into S0 galaxies (Moran, Ellis \\& Treu 2006). Indeed, Goto et al. (2003b) found that the Petrosian radius of S0s is smaller than that of spirals, which is consistent with the above scenario. Although these studies have provided significant broad implications concerning the origin of passive spiral galaxies, the details of the underlying physical mechanisms, especially the time-scale on which in-falling cluster galaxies terminate their star formation, are still uncertain. Various mechanisms with different time-scales have been proposed. Because star formation could be sustained by cold gas accreted onto the disc, the time-scale is closely related to the time at which the reservoir of cold gas is removed by means of galaxy-intracluster medium (ICM) interactions (Treu et al. 2003; Diaferio et al. 2001; Kauffmann et al. 1999). A possible mechanism that occurs on a relatively short time-scale is ram-pressure stripping (Gunn \\& Gott 1972; Fujita 2001; Fujita \\& Nagashima 1999; Fujita \\& Goto 2004; Abadi, Moore \\& Bower 1999; Quilis, Moore \\& Bower 2000). When a galaxy falls into a dense region of the cluster, ram pressure caused by the motion of the galaxy relative to the dense ICM removes the cold interstellar gas in the disc that is the fuel for the star formation (Gunn \\& Gott 1972). Vollmer et al. (2006) modelled the heavily ram-pressure-stripped galaxy NGC 4522 in the Virgo Cluster (Kenney, van Gorkom \\& Vollmer 2004). The ram-pressure-stripping model successfully reproduces the observed H I gas deficiency and the truncated gas disc of the galaxy (Vollmer et al. 2006). The time-scale for the ram-pressure stripping to terminate star formation was estimated to be 0.01-0.1 Gyr from numerical simulations (Abadi et al. 1999; Quilis et al. 2000; Fujita \\& Nagashima 1999). Some authors, however, argue that ram pressure alone cannot explain the observed decline in star formation in cluster galaxies (Treu et al. 2003; Balogh et al. 2002; Kodama \\& Bower 2001). Treu et al. (2003) reported the mild decline in star formation at the periphery of the cluster Cl 0024$+$16, where the ram pressure may not be effective in stripping cold disc gas. Other possible mechanisms have been proposed that require relatively longer time-scales to strip the galactic halo gas (Larson, Tinsley \\& Caldwell 1980), and these have been termed 'strangulation' (Fujita 2004; Tanaka et al. 2004) or 'starvation' (Treu et al. 2003; Boselli et al. 2006; Bekki et al. 2002). 'Strangulation' involves the stripping of warm halo gas through the interaction with the ICM in a situation where further supply of halo gases to the disc is disrupted. Because this can occur even in less dense environments, this may explain the observed decline of star formation among galaxies in the outer regions of clusters. Mechanisms such as mergers cannot be responsible for creating passive spiral galaxies because they may disturb the spiral arms. Similarly, 'harassment', high-speed gravitational encounters between galaxies (Moore et al. 1996), may also lead to changes in morphology and thus cannot be a dominant mechanism for transforming normal spiral galaxies into passive ones. In this paper, we perform spatially resolved spectroscopy on four of the passive spiral galaxies identified in our previous paper (Goto et al. 2003b) in order to obtain further constraints on the mechanism responsible for halting the star formation. The strength of the H$\\delta$ absorption line and 4000-{\\AA} break are compared with the simple stellar population (SSP) model constructed by Bruzual \\& Charlot (2003) in order to estimate the age of the stellar population. We then attempt to estimate the time-scale for the mechanism to create passive spiral galaxies. The method of data reduction and analysis are described in Section 2, the results are in Section 3, the discussion based on the comparison with the SSP model is presented in Section 4, and conclusions are given in Section 5. Unless otherwise stated, we adopt the best-fitting WMAP cosmology: (h, $\\Omega_{m}$, $\\Omega_{L}$) = (0.71, 0.27, 0.73) (Bennett et al. 2003). \\begin{table*} \\centering \\begin{minipage}{140mm} \\label{tab1:1} \\begin{tabular}{@{}lllcccccc@{}} \\hline & & & & & & \\multicolumn{2}{c}{Signal-to-noise ratio\\footnote{Signal-to-noise ratio measured by taking the standard deviation of the continuum waveband defined for spectral indices}}\\\\ Name & Ra(J2000)&Dec(J2000)&$z$&$\\sigma_{V}$ km s$^{-1}$\\footnote{Velocity dispersion $\\sigma_{V}$ and Petrosian radius $r_{\\rm P}$ in $r'$ band are taken from SDSS Data Release 5 \\citep{b3,Adelman} }&$r_{\\rm P}$ arcsec & ${\\rm AP}_{in}$&${\\rm AP}_{out}$ & $M_r$\\footnote{Absolute magnitude in $r'$ band}\\\\ \\hline SDSS J021534.35$-$090537.0 & 02 15 34.35&$-$09 05 37.0&0.0687 & 112$\\pm$7&15.62 &19.7& 9.8& $-$22.58\\\\ SDSS J024732.02$-$065137.5 & 02 47 32.02&$-$06 51 37.5&0.0705 &131$\\pm$9 &10.31 & 17.2&12.0& $-$21.75\\\\ SDSS J033322.66$-$000907.5&03 33 22.66&$-$00 09 07.5& 0.0838 & - & 6.615 & 11.1 & 7.0 & $-$20.89\\\\ SDSS J074452.51$+$373852.7&07 44 52.51&$+$37 38 52.7& 0.0743 &94$\\pm$9& 8.329 & 9.4 & 4.6 & $-$20.80\\\\ \\hline \\end{tabular} \\caption{Properties of target galaxies. Aperture radii for the spatially resolved analysis ($r_{\\rm P}$) are also shown.} \\end{minipage} \\end{table*} \\begin{figure*} \\begin{minipage}{140mm} \\includegraphics[width=60mm,clip]{033.eps} \\includegraphics[width=60mm,clip]{074.eps} \\caption{SDSS $g', r', i'$ composite images \\citep{Fukugita} of target galaxies (left: SDSS J033322.65$-$000907.5, right: SDSS J074452.52$+$373852.7)} \\label{SDSSpictures} \\end{minipage} \\end{figure*} \\section[]{The Method} \\label{sec:section2} \\subsection{Sample selection} The target galaxies, SDSS J021534.35$-$090537, J024732.02$-$065137.5, J033322.65$-$000907.5 (the left panel of Fig. 1) and J074452.52$+$373852.7 (the right panel of Fig. 1), are a subset of the passive spiral galaxies selected in Goto et al. (2003b) from the volume-limited sample of the SDSS data. Galaxies are selected based on the following criteria: (1) the inverse concentration parameter, which is defined as the ratio of the Petrosian 50 per cent radius to the Petrosian 90 per cent radius, is less than 0.5 (details of the use of this parameter for classifying morphological type are given in Shimasaku et al. 2001) in order to select galaxies having spiral morphology; (2) the absence of [O II] and H$\\alpha$ emission lines (the measured values of equivalent width are less than 1 $\\sigma$ error, Goto et al. 2003b), which are indicative of ongoing star formation activity. \\subsection{Observation and Data reduction} \\label{subsec:2} The observations were carried out using the Dual Imaging Spectrograph (DIS) installed on the Apache Point Observatory 3.5-m telescope on 2004 October 18. Both the blue and red cameras on the DIS were used in a medium-dispersion mode, with the dispersion covering the wavelength range 3000-9000 {\\AA}. The pixel scales of the spatial axis for the blue and the red cameras were 0.42 and 0.40 arcsec pix$^{-1}$, respectively. To perform spatially resolved spectroscopy, a long slit with a slit width of 1.5 arcsec was used. Each sample galaxy was observed three times with an exposure time of 1000-1500 s. After bias-subtraction and flat-fielding had been applied, the three frames were combined into one frame. Spectra of standard stars, HR 718 and Hilt 600, were taken with an exposure time of 20 and 1 s, respectively, and used for flux calibration. The seeing size measured using the full width at half maximum (FWHM) of the point spread function of Hilt 600, which was monitored during the observation of the target galaxies, was $\\sim$1.7 and $\\sim$1.3 arcsec for the red and the blue camera, respectively. To perform the spatially resolved analysis, we refer to the Petrosian radius in r' band ($r_{\\rm P}$), which is a measure of the surface brightness profile of galaxies, obtained through the SDSS (Blanton et al. 2001). The values are presented in Table 1. The long-slit data were divided into 11 spatial bins using the iraf routine apall, which outputs 11 spectra for all bins. Then, the three bins around the centre and the remaining eight bins sampling the spectra of two sides of the galaxies were summed, respectively, to increase the signal-to-noise ratio. Hereafter, these summed data for the inner and outer parts of the galaxies are designated as AP$_{in}$ and AP$_{out}$, respectively. AP$_{in}$ samples approximately $r$ $\\leq$ 0.27$r_{\\rm P}$, whereas AP$_{out}$ samples 0.27$r_{\\rm P}$ $<$ r $\\leq$ 1$r_{\\rm P}$, where $r$ represents the distance from the galaxy centre. The diameter of the inner part, $\\sim$3.7 arcsec, is much larger than the seeing size ($<$1.7 arcsec). Wavelength calibration was performed using observations of a HeNeAr lamp. The sensitivity function was obtained using the spectra of the standard stars HR 718 and Hilt 600. Because it was difficult to fit the whole wavelength range with a single sensitivity function, we excluded the data points at the edges of each red and blue spectrum. The resulting spectral coverage, included in the fitting of the sensitivity correction, is 3700-5650 {\\AA} for the blue spectrum and 6300-9500 {\\AA} for the red spectrum. The wavelength resolution was measured using the FWHM of the emission lines of the HeNeAr lamp spectra. The result was $\\sim 8.5 \\pm 0.5$ {\\AA} FWHM at approximately 5000 {\\AA}. \\subsection{Measurement of spectral features} \\label{subsec:3} \\begin{figure*} \\begin{center} \\label{fig:ha} \\includegraphics[width=45mm,angle=270]{ha_021.ps} \\includegraphics[width=45mm,angle=270]{ha_024.ps} \\includegraphics[width=45mm,angle=270]{ha_033.ps} \\includegraphics[width=45mm,angle=270]{ha_074.ps} \\caption{The rest-frame spectra around the ${\\rm H}\\alpha$ emission line for the observed galaxies. Solid lines show the spectra for ${\\rm AP}_{out}$ and dotted lines show the spectra for ${\\rm AP}_{in}$. } \\end{center} \\end{figure*} Before we measured the spectral features, the redshift for each galaxy was computed from the Ca H line at 3970 {\\AA}, one of the most prominent features in the obtained spectra, by fitting a Gaussian for this line with the iraf splot routine. The wavelengths of the observed spectrum were shifted to the rest-frame based on the derived redshifts. The derived redshifts are shown in Table 1. The Lick/IDS absorption-line index system (Trager et al. 1998; Worthey 1994) was used to measure the strengths of absorption features in the spectra. The wavelength definitions for the index measurement were taken from Worthey \\& Ottaviani (1997) for the H$\\gamma$ and H$\\delta$ indices, and from Trager et al. (1998) for the other 21 indices. The equivalent width or magnitudes for each index were computed as defined in Trager et al. (1998). We also measured the 4000-{\\AA} break (D$_{4000}$) to estimate the age of the stellar population, as it is a broad feature and can be measured with a greater signal-to-noise ratio than can individual lines. D$_{4000}$ is widely used as a diagnostic of the age and metallicity of stellar populations and can be compared with values in the literature (Bruzual 1983; Gorgas et al. 1999; Kauffmann et al. 2003). D$_{4000}$ is calculated as defined in Balogh et al. (1999), who take a narrower spectral region for the red and blue continua than do Bruzual (1983). The reason for using this narrower definition is that it is less affected by the uncertainty in the sensitivity correction, and, more importantly, by the reddening effect (Balogh et al. 1999). In order to confirm the absence of star formation activity for the observed candidates of passive spiral galaxies, equivalent widths of H$\\alpha$ and [O II] emission line were measured. Fig. 2 shows the spectral regions around the H$\\alpha$ emission lines. Two of the four target galaxies (SDSS J021534.35$-$090537 and SDSS J024732.02$-$065137.5) are found to show detectable H$\\alpha$ emission (H$\\alpha$ $-$ 1$\\sigma$ error $\\geq$ 10 {\\AA}) at AP$_{out}$, and thus do not meet the criteria for passive spiral galaxies defined in Goto et al. (2003b). The [O II] emission lines, whose equivalent widths are less than 7 {\\AA}, were also detected at AP out, indicating the presence of current star formation activity in the exterior regions of these galaxies. The non-detection of [O II] in the SDSS data was presumably the result of the aperture of the SDSS spectroscopic fiber (diameter of 3 arcsec), which samples only the inner parts of the galaxies, where the observed light may be dominated by the bulge component over the disc component (Abazajian et al. 2005). The results indicate the importance of using the whole light, including that from the outer regions of the galaxies, when identifying and investigating passive spiral galaxies. We restrict our discussion to galaxies with no prominent star formation activity over the whole galaxy. The remaining two galaxies, SDSS J033322.65$-$000907.5 and SDSS J074452.52$+$373852.7, are hereafter denoted as SDSSJ0333$-$0009 and SDSSJ0744$+$3738. The velocity dispersion (available only for SDSSJ0744$+$3738), Petrosian radius, absolute magnitude in the r' band obtained by the SDSS, and the measured values of redshift and signal-to-noise ratio are shown in Table 1. The g', r', i'-composite SDSS images (Fukugita et al. 1996) of SDSSJ0333$-$0009 and SDSSJ0744$+$3738 are shown in Fig. 1. ", "conclusions": "\\label{sec:section4} In the following subsection, we first try to estimate the metallicity and the effects of $\\alpha$-enhancement on the absorption strengths for the observed galaxies using the measured Lick indices. The super-solar $\\alpha$/Fe ratio could affect the strength of the H$\\delta_{\\rm A}$ and the H$\\delta_{\\rm F}$ indices, as discussed in Thomas, Maraston \\& Bender (2003), which leads to an underestimate of the age of stellar populations. Considering the estimated metallicity and the effect of $\\alpha$-enhancement, Section 4.2 discusses the light-averaged ages of the stellar populations using a H$\\delta_{\\rm F}$-D$_{4000}$ plane. The Lick indices measured with $>$1$\\sigma$ for both the inner and outer regions were Fe4383, Fe4668, H$\\beta$, Fe5270 and H$\\delta_{\\rm F}$ for SDSSJ0333$-$0009. For SDSSJ0744$+$3738, G4300, Fe4531 and Fe4668 were detected with $>$1$\\sigma$. Unfortunately, the H$\\delta_{\\rm F}$ index, which is used as an age indicator, does not have a good enough signal-to-noise ratio for SDSSJ0744$+$3738. Furthermore, neither the inner nor the outer region of this galaxy has an Mgb index, which is used as an indicator of metallicity and $\\alpha$-enhancement, with a signal-to-noise ratio $>$1. We therefore focus our discussion of the metallicity and $\\alpha$-enhancement on SDSSJ0333$-$0009, for which the iron indices and the Mgb indices were measured with a relatively high signal-to-noise ratio. \\subsection{Metallicity and $\\alpha$-enhancement} \\label{sec:metal} The measured Lick indices (Fe4383, Fe4668, H$\\beta$, Fe5270 and H$\\delta_{\\rm F}$) for SDSSJ0333$-$0009 indicate that the inner region is more metal-rich than the outer region. Fig. 6 shows the strength of these indices for AP$_{in}$ (open triangles) and AP$_{out}$ (filled triangles) as a function of the Mgb index. The SSP models of Thomas et al. (2003) with age = 1.0, 5.0, 10.0, 15.0 Gyr and metallicity Z = 0.67, 0.00(Z$_{\\odot}$), -2.25, assuming [$\\alpha$/Fe] = 0.0, are overlaid. A comparison of the strength of the observed indices with that of the SSP models suggests that the observed indices for both the inner and outer regions are consistent with the model with super-solar metallicity. The iron indices (Fe4383, Fe4668 and Fe5270), especially, suggest that there is a radial gradient in the metallicity; the observed indices for the inner region are marginally consistent with [Z/H] = 0.67 models, whereas for the outer region the observed indices are reproduced by the model with [Z/H] = 0.00 in the Fe4668-Mgb plane and [Z/H] = 0.00-0.67 in the Fe5270-Mgb plane. Therefore, although the index strengths are associated with large uncertainties, they suggest that the inner region is more metal-rich than the outer region. It should be noted that H$\\delta_{\\rm F}$ for both regions and Fe4383 for the outer region were not reproduced by the models for [$\\alpha$/Fe] = 0.0. Rather, they seem to be consistent with models with [$\\alpha$/Fe] = 0.5, as discussed in the next section. Comparisons of the Fe4383 indices, which are especially sensitive to the $\\alpha$-enhancement (Thomas et al. 2003), with the model suggest that the outer region of this galaxy possibly has a super-solar $\\alpha$/Fe ratio. Fig. 7 shows the strength of the Fe4383 index as a function of the Mgb index. Models with [$\\alpha$/Fe] = 0.0 and 0.5 for ages and metallicities the same as those shown in Fig. 6 are overlaid. Open and filled triangles show the observed index strength for AP$_{in}$ and AP$_{out}$, respectively. This figure indicates that the strength of Fe4383 for AP$_{out}$ is more consistent with the models with [$\\alpha$/Fe] = 0.5 than it is with those with [$\\alpha$/Fe] = 0.0. In order to elucidate the dependence of the H$\\delta$ absorption strength on [$\\alpha$/Fe], Fig. 8 shows the model predictions for the H$\\delta_{\\rm F}$ index as a function of the Mgb index for [$\\alpha$/Fe] = 0.0 and 0.5. This illustrates that the difference in H$\\delta_{\\rm F}$ absorption owing to the $\\alpha$-enhancement is expected to be $\\sim$ 0.8-1.2 {\\AA} , depending on the age and metallicity of the stellar population. The inner region (open triangles) is roughly consistent with the model with age $\\sim$ 1.0-1.5 Gyr, metallicity Z $\\sim$ 0.67, and [$\\alpha$/Fe] = 0.5. The outer region (filled triangles), by contrast, shows stronger H$\\delta_{\\rm F}$ absorption, which suggests that the stellar population in the outer region is, on average, younger than that in the inner region. The age and the metallicity implied from the plot for H$\\delta_{\\rm F}$ against Mgb are $<$ 1.0 Gyr and Z $\\sim$ 0.00-0.67, respectively. Although the metallicity may be difficult to determine quantitatively from Fig. 8 because of the quite large uncertainty in the Mgb index, Z $\\sim$ 0.00-0.67 is consistent with the value implied from the strength of the Fe4668 and Fe5270 indices against the Mgb index as shown in Fig. 6. \\begin{figure*} \\begin{center} \\begin{tabular}{ccc} \\rotatebox{270}{\\resizebox{38mm}{!}{\\includegraphics{mgb_metals_SDSS033322_Fe4383.ps}}} & \\rotatebox{270}{\\resizebox{38mm}{!}{\\includegraphics{mgb_metals_SDSS033322_Fe4668.ps}}} & \\rotatebox{270}{\\resizebox{38mm}{!}{\\includegraphics{mgb_metals_SDSS033322_Hb.ps}}} \\\\ \\rotatebox{270}{\\resizebox{38mm}{!}{\\includegraphics{mgb_metals_SDSS033322_Fe5270.ps}}} & \\rotatebox{270}{\\resizebox{38mm}{!}{\\includegraphics{mgb_metals_SDSS033322_HdF.ps}}} & \\\\ \\end{tabular} \\caption{he Lick indices measured with $>$1$\\sigma$ as a function of the Mgb index for SDSSJ0333$-$0009. Open and filled triangles show the measured index strength for AP$_{in}$ and AP$_{out}$, respectively. The solid lines show the models with [Z/H] = $-2.25$, $0.00$(Z$_{\\odot}$) and 0.67 (from left to right). The dotted lines show the models with ages = 1, 5, 10 and 15 Gyr (for the Fe4383, Fe4668 and Fe5270 indices, the age increases from left to right; for the H$\\beta$ and H$\\delta_{\\rm F}$ indices, the age increase, from top to bottom).} \\label{mgb_metals} \\end{center} \\end{figure*} \\begin{figure*} \\begin{tabular}{c@{\\hspace{1cm}}c} \\begin{minipage}{0.45\\hsize} \\begin{center} \\includegraphics[width=55mm,angle=270]{mgb_fe4383.ps} \\end{center} \\caption{The Fe4383 index for AP$_{in}$ and AP$_{out}$ as a function of the Mgb index. The models with [$\\alpha$/H] = 0.0 and 0.5 are overlaid. For both values of [$\\alpha$/H], the solid lines show the models with [Z/H] = $-2.25$, 0.00 and 0.67 (from bottom to top) and the dotted lines show the models with ages = 1, 5, 10 and 15 Gyr (from bottom to top)} \\label{mgb_fe4383} \\end{minipage} & \\begin{minipage}{0.45\\hsize} \\begin{center} \\includegraphics[width=55mm,angle=270]{mgb_HdF.ps} \\end{center} \\caption{The H$\\delta_{\\rm F}$ index for AP$_{in}$ and AP$_{out}$ as a function of the Mgb index. The models with [$\\alpha$/H] = 0.0 and 0.5 are overlaid. For both values of [$\\alpha$/H], the solid lines show the models with [Z/H] = $-2.25$, 0.00 and 0.67 (from left to right) and the dotted lines show the models with ages = 1.0, 5.0, 10.0 and 15.0 Gyr (from bottom to top).} \\label{mgb_hdF} \\end{minipage} \\end{tabular} \\end{figure*} \\subsection{Comparison with the SSP models} \\label{sec:age} In the following discussions, we use the H$\\delta_{\\rm F}$ index and the D$_{4000}$ feature as age indicators of the stellar population. The H$\\delta_{\\rm F}$-D$_{4000}$ plot is widely used as a diagnostic tool to estimate the age of a stellar population in galaxies (Kauffmann et al. 2003; Poggianti \\& Barbaro 1997; Balogh et al. 1999). The great advantage in using the H$\\delta_{\\rm F}$ and D$_{4000}$ indices as age indicators is that they are less affected by the reddening caused by interstellar dust (Kauffmann et al. 2003). We note that the ages discussed in this section refer to the light-averaged ages of stellar populations, and thus we aimed to estimate the time at which star formation ceased in the samples qualitatively. In Fig. 9, the measured H$\\delta_{\\rm F}$ index and the values of D$_{4000}$ are compared with predictions from the simple stellar population (SSP) models in the galaxev package of Bruzual \\& Charlot (2003). We used the model with the Salpeter initial mass function with lower and upper cut-offs of stellar masses of 0.1 ${\\rm M}_{\\odot}$ and 100 ${\\rm M}_{\\odot}$, respectively. The model assumes an instantaneous burst of star formation at age = 0 for the various metallicities (Bruzual \\& Charlot 2003). The spectra of the SSP model used in Fig. 9 were broadened to the instrumental resolution of the observations (8.5 {\\AA} FWHM), which corresponds to a velocity dispersion of $\\sigma$ $\\sim$ 197 km s$^{-1}$. The spectral broadening was applied using the program vel\\_disp included in the galaxev package (Bruzual \\& Charlot 2003). The resulting SSP model tracks for metallicities [Fe/H] = $-1.6464$, $-0.6392$, $+0.0932$, $+0.5595$ are overlaid in Fig. 9. Points with the same ages (1.0, 1.6, 2.5, 4.0, 6.2, 10.0 Gyr) are joined with dotted lines. Based on the Fe4383 index, the outer regions of SDSSJ0333$-$0009 have [$\\alpha$/Fe] $>$ 0.0, as discussed in the previous subsection. Because a stellar population which has [$\\alpha$/Fe] $>$ 0.0 is reported to be enhanced in H$\\delta$ absorption strength compared to that having solar $\\alpha$/Fe ratios with similar age (Thomas et al. 2004), we have to take this effect into account to estimate the stellar population ages using H$\\delta_{\\rm F}$ indices. According to the stellar population model of Thomas et al. (2003), the enhanced strength in H$\\delta_{F}$ indices is up to $\\sim$ 1.2 {\\AA} for the model with [$\\alpha$/Fe] = 0.5. Therefore, the data points should be shifted downwards from those actually plotted in Fig. 6 by up to $\\sim$ 1.2 {\\AA} following the correction for the $\\alpha$-enhancement. For SDSSJ0744$-$3738, [$\\alpha$/Fe] could not be determined from the observed spectra, because the Mg indices were not measured with a high enough signal-to-noise ratio. Therefore, the additional uncertainty arising from the $\\alpha$-enhancement should be taken into account when considering H$\\delta_{\\rm F}$ versus D$_{4000}$ diagnostics. Another point of caution in using the H$\\delta_{\\rm F}$-D$_{4000}$ plot is that the stellar absorption of H$\\delta_{\\rm F}$ could be contaminated by nebular emission (Kauffmann et al. 2003; Goto et al. 2003a). The nebular emission of H$\\alpha$ is weak for SDSSJ0333$-$0009 and SDSSJ0744$+$3738 ($<-4$ {\\AA}), and, furthermore, H$\\beta$ was detected in absorption. The effect of emission filling on the H$\\delta_{\\rm F}$ absorption is expected to be negligible ($\\sim$ 0.3 {\\AA} assuming the Case B recombination value for the H$\\alpha$/H$\\delta$ nebulae emission line ratio without dust extinction). In the following subsection, we estimate the light-averaged age of the stellar populations in the inner and outer regions for the observed galaxies. \\subsubsection{SDSS J033322.65$-$000907.5} The measured H$\\delta_{\\rm F}$ and D$_{4000}$ for AP$_{in}$ and AP$_{out}$ are shown in Fig. 9 by the open and filled triangles, respectively. If we assume a metallicity [Fe/H] $> + 0.0932$ and a correction for the $\\alpha$-enhancement of $\\sim$1.2 {\\AA} as suggested in Section 4.1, the data for APout are broadly consistent with a model with age older than $\\sim$ 1.0 Gyr. By contrast, for the inner region (AP$_{in}$) the observed H$\\delta_{\\rm F}$ index seems stronger than the value that can be reproduced by the model with super-solar metallicity ([Fe/H] = $+0.5595$). However, the observed strength of D$_{4000}$ for AP$_{in}$ suggests an age older than $\\sim$ 1.6 Gyr. Overall, the H$\\delta_{\\rm F}$-D$_{4000}$ plane suggests that the average age of the stellar population is younger in the outer region than in the inner region for this galaxy. \\subsubsection{SDSS J074452.52$+$373852.7} The open and filled squares in \\ref{fig:4} correspond to the data for AP$_{in}$ and AP$_{out}$, respectively, for SDSSJ0744$+$3738. For this galaxy, the metallicity and [$\\alpha$/Fe] are not available because of the large errors in the Lick indices. Although the determination of the age of the stellar population for SDSSJ0744$+$3738 is quite uncertain, the observed large difference in D$_{4000}$ between the core and the exterior of the galaxy imply that the two regions have distinct stellar populations in terms of their age and/or metallicity. For APin, the large value of D$_{4000}$ is marginally consistent with the model with metallicity [Fe/H] = $+0.5595$ in the H$\\delta_{\\rm F}$-D$_{4000}$ plane. Furthermore, in the outer region (AP$_{out}$), the H$\\delta_{\\rm F}$ index is quite strong (5.5$\\pm$11.1{\\AA}), although we note that the index strength is associated with large errors. Galaxies with strong H$\\delta$ absorption are classified as 'H$\\delta$-strong' (HDS) galaxies \\citep{b57,b20,b9,2004MNRAS.348..515G,b61}. As strong H$\\delta$ absorption could arise from A-type stars in the main-sequence phase \\citep{b2}, HDS galaxies probably terminated their starburst activity $\\sim$0.1-1.5 Gyr ago \\citep{b32,b9}. Therefore, the strong H$\\delta_{\\rm F}$ absorption in the outer regions of the observed galaxies could be clear evidence that they terminated their star formation a few Gyrs ago. Nevertheless, because of the large uncertainty associated with H$\\delta_{F}$, we need further observations with higher signal-to-noise ratios to detect a clear signature of quenched star formation. \\begin{figure} \\begin{center} \\includegraphics[width=80mm]{inst_burst.ps} \\caption{Equivalent width of H$\\delta_{\\rm F}$ as a function of D$_{4000}$. The open and filled triangles represent the data for AP$_{in}$ and AP$_{out}$, respectively, for SDSSJ0333$-$0009. The open and filled squares represent the data for APin and APout, respectively, for SDSSJ0744$+$3738. Overlaid lines show predictions of SSP models broadened to the instrumental resolution ($\\sigma$ = 197 km s$^{-1}$) with various metallicities [Fe/H] = $-1.6464$, $-0.6392$, $+0.0932$, $+0.5595$ (Bruzual \\& Charlot 2003). Dotted lines connect points of the same age (from bottom to top, 10.0, 6.2, 4.0, 2.5, 1.6 and 1.0 Gyr, respectively).} \\label{fig:4} \\end{center} \\end{figure} \\subsubsection{Implications} It is interesting that through the spectral diagnostics of spatially separated regions of passive spiral galaxies we can estimate the history of these galaxies. Even our spectra with relatively low signal-to-noise ratios suggest that passive spiral galaxies were in a star-forming phase for several gigayears before ceasing activity. Moreover, one of the passive spirals, SDSSJ0744$+$3738, probably terminated its star formation approximately 1-2 Gyr ago. If confirmed, this result will provide us with an important constraint in uncovering the underlying (cluster-related) physical mechanism responsible for the creation of passive spiral galaxies; for example, ram-pressure stripping \\citep{b24,b37, b38,b47,b48,b25} stops star formation much quicker than the strangulation scenario \\citep{b52,b46}. To reach a firm conclusion on the subject, however, further observations are needed: it is well known that the age/metallicity estimates are degenerate, for example the spectra of an old ($>$ 2 Gyr) stellar population looks almost identical when the age is doubled and the total metallicity reduced by a factor of 3 \\citep{b62}. It is therefore important to break this degeneracy by obtaining spectra at a higher signal-to-noise ratio to measure indices sensitive only to age \\citep[e.g. Balmer lines;][]{b4} and only to metallicity \\citep[e.g.][]{b19}. Furthermore, we only had enough observing time for two passive spiral galaxies. As larger samples of passive spiral galaxies are now becoming available \\citep[e.g.][]{b1}, it is important to investigate the statistical proportion of passive spiral galaxies that do not suffer from shot-noise to draw firm conclusions on this subject." }, "0801/0801.0167.txt": { "abstract": "Particle physics candidates for cosmological dark matter are usually considered as neutral and weakly interacting. However stable charged leptons and quarks can also exist and, hidden in elusive atoms, play the role of dark matter. The necessary condition for such scenario is absence of stable particles with charge -1 and effective mechanism for suppression of free positively charged heavy species. These conditions are realized in several recently developed scenarios. In scenario based on Walking Technicolor model excess of stable particles with charge -2 and the corresponding dark matter density is naturally related with the value and sign of cosmological baryon asymmetry. The excessive charged particles are bound with primordial helium in techni-O-helium \"atoms\", maintaining specific nuclear-interacting form of dark matter. Some properties of techni-O-helium Universe are discussed. {\\bigbreak\\bgroup\\narrower\\small \\baselineskip=10pt plus .2pt \\parindent 0pt} \\def\\endabstract{\\par\\egroup\\bigbreak} % --------------------- References ----------------------- % Environnement r\u00e9f\u00e9rences \\newenvironment{fref}{ \\renewcommand\\refname{R\\'ef\\'erences} \\begin{thebibliography}{99} \\setlength\\itemsep {-3pt} \\small}{\\end{thebibliography}} % id en anglais \\newenvironment{eref}{ \\renewcommand\\refname{References} \\begin{thebibliography}{99} \\setlength\\itemsep {-3pt} \\small}{\\end{thebibliography}} % ------------------ Date reception ----------------------- % \\outer\\def\\man#1{\\medskip \\noindent{\\it (Manuscrit re\\c cu le #1)}} % % Macros possibles \\def\\eqalign#1{\\setlength{\\arraycolsep}{.2em} \\renewcommand{\\arraystretch}{1.5} \\begin{array}{rl} #1\\end{array}} %-------------------- Figures ----------------------------- \\long\\def\\Fig#1{{\\small \\noindent{\\bf Figure #1. }\\par}} %\\makeatother \u0000\u00faS\u0001\u0088XJ\u0003%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u0000\u000f\u0000\u0001\u000e&\u008e\u0010\b\\bibitem{BKSR} Belotsky, K.M., Fargion, % A adapter \\`a l'ordinateur, ici pour un Macintosh %\\usepackage[T1]{fontenc} \\usepackage[applemac]{inputenc} \\usepackage{amssymb,amsmath} \\def\\volnum{Manuscrit} %\\def\\volnum{Volume xx no y, 200z} % pour la c\\'esure %\\french % Format des AFLB \\usepackage{aflbcours} \\pdebut{1} \\def\\pacs#1{\\LP P.A.C.S.: #1} \\def\\s{{\\,\\rm s}} \\def\\g{{\\,\\rm g}} \\def\\eV{\\,{\\rm eV}} \\def\\keV{\\,{\\rm keV}} \\def\\MeV{\\,{\\rm MeV}} \\def\\GeV{\\,{\\rm GeV}} \\def\\TeV{\\,{\\rm TeV}} \\def\\sv{\\left<\\sigma v\\right>} \\def\\({\\left(} \\def\\){\\right)} \\def\\cm{{\\,\\rm cm}} \\def\\K{{\\,\\rm K}}\\title{Dark matter from stable charged particles?} \\titleshort{Dark matter from stable charged particles \\dots} \\author{Maxim Yu. Khlopov} \\authorshort{Maxim Khlopov} \\address{Centre for Cosmoparticle physics \"Cosmion\", \\\\ Miusskaya Pl. 4, 125047 Moscow, Russia\\\\ Moscow Engineering Physics Institute, 115409 Moscow, Russia, and \\\\ APC laboratory 10, rue Alice Domon et L\u00e9onie Duquet 75205 Paris Cedex 13, France\\\\Maxim.Khlopov@roma1.infn.it} \\setlength{\\arraycolsep}{0pt} \\begin{document} A widely accepted viewpoint is that the approriate candidates for cosmological dark matter are neutral and weakly interacting particles. However it turns out that even stable charged leptons and quarks can exist and, hidden in elusive atoms, play the role of dark matter. The necessary condition for such scenario is absence of stable particles with charge -1 and effective mechanism of suppression for free positively charged heavy species. These conditions are realised in several recently developed scenarios. In scenario based on Walking Technicolor model excess of stable particles with charge -2 (without stable particles with charge -1) is naturally related with a cosmological baryon excess. ", "introduction": "The modern theory of Universe, based on General Relativity, has evolved the triumph of Einstein's ideas by putting cosmological term, first introduced by A.Einstein in 1917 \\cite{Lambda}, in the \"standard\" $\\Lambda$CDM model. The corresponding dark energy is the dominant element of the modern Universe, maintaining $70\\%$ of its total density. General Relativity and Dark Energy maintain the frame for the portrait of elementary particles in the Universe. To survive in the Universe the particles should be stable, as are nuclei and electrons, composing the visible matter. However, one must also explain the modern dark matter density, corresponding to $25\\%$ of total density and exceeding the baryonic matter density by a factor of 5. The widely shared belief is that dark matter is nonbaryonic and consists of new stable particles. For a particle with the mass $m$ the particle physics time scale is $t \\sim 1/m$ (here and further, if not indicated otherwise, we use the units $\\hbar = c = k = 1$), so in particle world we refer to particles with lifetime $\\tau \\gg 1/m$ as to metastable. To be of cosmological significance metastable particle should survive after the temperature of the Universe $T$ fell down below $T \\sim m$, what means that the particle lifetime should exceed $t \\sim (m_{Pl}/m) \\cdot (1/m)$. Such a long lifetime should find reason in the existence of an (approximate) symmetry. From this viewpoint, cosmology of dark matter is sensitive to the most fundamental properties of microworld, to the conservation laws reflecting strict or nearly strict symmetries of particle theory. One can formulate the set of conditions under which new particles can be considered as candidates to dark matter (see e.g. \\cite{book,Cosmoarcheology,Bled06,Bled07} for review and reference). \\begin{itemize} \\item[$\\bullet$] The particles should be stable or have lifetime larger, than age of the Universe. \\item[$\\bullet$] They should fit the measured density of dark matter. Effects of their decay or annihilation should be compatible with the observed fluxes of electromagnetic background radiation and cosmic rays. \\item[$\\bullet$] More complicated forms of scalar fields, primordial black holes and even evolved primordial large scale structures are also possible, but in the latter case the contribution to the total density is restricted by the condition of the observed homogeneity and isotropy of the Universe\\item[$\\bullet$] The candidates for dark matter should decouple from plasma and radiation at least before the beginning of matter dominated stage. This is necessary to provide formation of large scale structure at the observed level of anisotropy of the cosmic microwave background radiation. \\end{itemize} The easiest way to satisfy these conditions is to involve neutral weakly interacting particles. However it is not the only particle physics solution for the dark matter problem. As we show here, new stable particles can have electric charge, but effectively behave as neutral and sufficiently weakly interacting. Recently several elementary particle frames for heavy stable charged particles were proposed: \\begin{itemize} \\item[(a)] A heavy quark and heavy neutral lepton (neutrino with mass above half the Z-Boson mass) of fourth generation \\cite{N,Legonkov}; which can avoid experimental constraints \\cite{Q,Okun} and form composite dark matter species \\cite{I,lom,KPS06,Khlopov:2006dk}; \\item[(b)] A Glashow's ``sinister'' heavy tera-quark $U$ and tera-electron $E$, which can form a tower of tera-hadronic and tera-atomic bound states with ``tera-helium atoms'' $(UUUEE)$ considered as dominant dark matter \\cite{Glashow,Fargion:2005xz}. \\item[(c)] AC-leptons, predicted in the extension \\cite{5} of standard model, based on the approach of almost-commutative geometry \\cite{bookAC}, can form evanescent AC-atoms, playing the role of dark matter \\cite{5,FKS,Khlopov:2006uv,Khlopov:2006dk}.\\end{itemize} Finally, it was recently shown in \\cite{KK} that an elegant solution is possible in the framework of walking technicolor models \\cite{Sannino:2004qp,Hong:2004td,Dietrich:2005jn,Dietrich:2005wk,Gudnason:2006ug,Gudnason:2006yj} and can be realized without an {\\it ad hoc} assumption on charged particle excess, made in the approaches (a)-(c). In all these models, predicting stable charged particles, the particles escape experimental discovery, because they are hidden in elusive atoms, maintaining dark matter of the modern Universe. It offers new solution for the physical nature of the cosmological dark matter. This approach differs from the idea of dark matter, composed of primordial bound systems of super heavy charged particles and antiparticles, proposed earlier to explain the origin of Ultra High Energy Cosmic Rays (UHECR) \\cite{UHECR}. To survive to the present time and to be simultaneously the source of UHECR super heavy particles should satisfy a set of constraints, which in particular exclude the possibility that they possess gauge charges of the standard model. The particles, considered here, participate in the standard model interactions and we discuss the problems, related with various scenarios of composite atom-like dark matter, formed by heavy electrically charged stable particles. \\vskip 30pt \\begin{fref} ", "conclusions": "Discussion} To conclude, the existence of heavy stable particles is one of the popular solutions for dark matter problem. If stable particles have electric charge, dark matter candidates can be atom-like states, in which negatively and positively charged particles are bound by Coulomb attraction. In this case there is a serious problem to prevent overproduction of accompanying anomalous forms of atomic matter. Indeed, recombination of charged species is never complete in the expanding Universe, and significant fraction of free charged particles should remain unbound. Free positively charged species behave as nuclei of anomalous isotopes, giving rise to a danger of their over-production. Moreover, as soon as $^4He$ is formed in Big Bang nucleosynthesis it captures all the free negatively charged heavy particles. If the charge of such particles is -1 (as it is the case for tera-electron in \\cite{Glashow}) positively charged ion $(^4He^{++}E^{-})^+$ puts Coulomb barrier for any successive decrease of abundance of species, over-polluting modern Universe by anomalous isotopes. It excludes the possibility of composite dark matter with $-1$ charged constituents and only $-2$ charged constituents avoid these troubles, being trapped by helium in neutral OLe-helium, O-helium (ANO-helium) or techni-O-helium states. The existence of $-2$ charged states and the absence of stable $-1$ charged constituents can take place in AC-model, in charge asymmetric model of 4th generation and walking technicolor model with stable doubly charged technibaryons and/or technileptons. To avoid overproduction of anomalous isotopes, an excess of $-2$ charged particles over their antiparticles should be generated in the Universe. In all the earlier realizations of composite dark matter scenario, this excess was put by hand to saturate the observed dark matter density. In walking technicolor model this abundance of -2 charged techibaryons and/or technileptons is connected naturally to the value and sign of cosmological baryon asymmetry. These doubly charged $A^{--}$ techniparticles bind with $^4He$ in the neutral techni-O-helium states. For reasonable values of the techniparticle mass, the amount of primordial $^4He$, bound in this atom like state is significant and should be taken into account in comparison to observations. A challenging problem is the nuclear transformations, catalyzed by techni-O-helium. The question about their consistency with observations remains open, since special nuclear physics analysis is needed to reveal what are the actual techni-O-helium effects in SBBN and in terrestrial matter. However, qualitatively one can expect a path for O-helium catalysis of heavy elements, making primordial heavy elements a signature for composite dark matter scenarios. The destruction of techni-O-helium by cosmic rays in the Galaxy releases free charged techniparticles, which can be accelerated and contribute to the flux of cosmic rays. In this context, the search for techniparticles at accelerators and in cosmic rays acquires the meaning of a crucial test for the existence of the basic components of the composite dark matter. At accelerators, techniparticles would look like stable doubly charged heavy leptons, while in cosmic rays, they represent a heavy $-2$ charge component with anomalously low ratio of electric charge to mass. The presented arguments enrich the class of possible particles, which can follow from extensions of the Standard Model and be considered as dark matter candidates. One can extend the generally accepted viewpoint that new stable particles should be neutral and weakly interacting. We have seen that they can also be charged and play the role of dark matter because they are hidden in atom-like states, which are not the source of visible light. The constraints on such particles are very strict and open a very narrow window for this new cosmologically interesting degree of freedom in particle theory. However, taking into account the exciting ability of O-helium to catalyze nuclear transformations of chemical elements, it is hard to estimate the expectation value of its discovery. To conclude, the existence of heavy stable charged particles can offer new solution for dark matter problem. Dark matter candidates can be atom-like states, in which negatively and positively stable charged particles are bound by Coulomb attraction. Primordial excess of these particles over their antiparticles implies the mechanism of its generation and is still an open problem for all the considered models. However, even if such mechanism exists, there is a serious problem of accompanying anomalous forms of atomic matter. Indeed, recombination of charged species is never complete in the expanding Universe, and significant fraction of free charged particles should remain unbound. Free positively charged species behave as nuclei of anomalous isotopes, giving rise to a danger of their over-production. Moreover, as soon as $^4He$ is formed in Big Bang nucleosynthesis it captures all the free negatively charged heavy particles. If the charge of such particles is -1 (as it is the case for tera-electron in \\cite{Glashow}) positively charged ion $(^4He^{++}E^{-})^+$ puts Coulomb barrier for any successive decrease of abundance of species, over-polluting modern Universe by anomalous isotopes. It excludes the possibility of composite dark matter with $-1$ charged constituents and only $-2$ charged constituents avoid these troubles, being trapped by helium in neutral OLe-helium or O-helium (ANO-helium) states. The existence of $-2$ charged states and the absence of stable $-1$ charged constituents can take place in AC-model and in charge asymmetric model of 4th generation. In the first case, pregalactic abundance of $C^{++}$ exceeds by ten orders of magnitude the terrestrial upper limit on anomalous helium and the mechanism of suppression of this abundance is inevitably accompanied by observable effects of recombination and implies the existence of $y$ charge, possessed by AC leptons. In the second case, owing to excess of $\\bar U$ anti-quarks primordial abundance of positively charged $U$-baryons is exponentially suppressed and anomalous isotope over-production is avoided. Excessive anti-$U$-quarks should retain dominantly in the form of anutium, which binds with excessive $\\bar N$ and then with $^4He$ in neutral ANO-helium. In the both cases, OLe-helium (or ANO-helium) should exist and its possible role in nuclear transformation is the serious danger (or exciting advantage?) for composite dark matter scenario. Galactic cosmic rays destroy ANO-helium (as well as OLe-helium), striking off $^4He$. It can lead to appearance of a free [anutium-$\\bar N$] component in cosmic rays, which can be as large as $[\\bar N \\Delta^{--}_{3 \\bar U}]/^4He \\sim 10^{-7}$ and accessible to PAMELA and AMS experiments. The estimation is two orders less in the case of free $A^{--}$ from cosmic ray destruction of OLe-helium \\cite{Khlopov:2006uv}. In the context of composite dark matter like \\cite{I,lom} or \\cite{FKS,Khlopov:2006uv} accelerator search for new stable quarks and leptons acquires the meaning of critical test for existence of its charged components. Such test will be possible in experiment ATLAS/LHC \\cite{lom,KPS06}. %\\Figure{crossec.eps}{Taken from \\cite{lom} cross sections of %production of heavy stable quarks and leptons in experiment ATLAS. %Dash line shows the lower (conservative) level of LHC sensitivity %expected for 1st year of its operation.}" }, "0801/0801.1564_arXiv.txt": { "abstract": "Observations show that the central black hole in galaxies has a mass $M$ only $\\sim 10^{-3}$ of the stellar bulge mass. Thus whatever process grows the black hole also promotes star formation with far higher efficiency. We interpret this in terms of the generic tendency of AGN accretion discs to become self--gravitating outside some small radius $R_{\\rm sg} \\sim 0.01 - 0.1$~pc from the black hole. We argue that mergers consist of sequences of such episodes, each limited by self--gravity to a mass $\\dme \\sim 10^{-3}M$, with angular momentum characteristic of the small part of the accretion flow which formed it. In this picture a major merger with $\\dmm \\sim M$ gives rise to a long series of low--mass accretion disc episodes, all chaotically oriented with respect to one another. Thus the angular momentum vector oscillates randomly during the accretion process, on mass scales $\\sim 10^3$ times smaller than the total mass accreted in a major merger event. We show that for essentially all AGN parameters, the disc produced by any accretion episode of this type has lower angular momentum than the hole, allowing stable co-- and counter--alignment of the discs through the Lense--Thirring effect. A sequence of randomly--oriented accretion episodes as envisaged above then produces accretion discs stably co-- or counter--aligned with the black hole spin with almost equal frequency. Accretion from these discs very rapidly adjusts the hole's spin parameter to average values $\\bar a \\sim 0.1-0.3$ (the precise range depending slightly on the disc vertical viscosity coefficient $\\alpha_2$) from any initial conditions, but with significant fluctuations ($\\Delta a\\sim \\pm 0.2$) about these. We conclude (a) AGN black holes should on average spin moderately, with the mean value $\\bar a$ decreasing slowly as the mass increases; (b) SMBH coalescences leave little long--term effect on $\\bar a$; (c) SNBH coalescence products in general have modest recoil velocities, so that there is little likelihood of their being ejected from the host galaxy; (d) black holes can grow even from stellar masses to $\\sim 5\\times 10^9\\, \\msun$ at high redshift $z\\sim 6$; (e) jets produced in successive accretion episodes can have similar directions, but after several episodes the jet direction deviates significantly. Rare examples of massive holes with larger spin parameters could result from prograde coalescences with SMBH of similar mass, and are most likely to be found in giant ellipticals. We compare these results with observation. ", "introduction": "Astronomers generally agree that the nuclei of most galaxies contain supermassive black holes (SMBH), but as yet have little clear idea of how they grow. Cosmological large--scale structure simulations strongly suggest that galaxy mergers are the basic motor of growth, and predict that a black hole typically acquires a mass $\\dmm$ of order its own mass $M$ in such mergers (e.g. di Matteo et al., 2005, Li et al., 2007, and references therein). This is reasonable, as growth rates of this order are needed in order to reach the hole masses observed in the nearby universe. If one assumes further that all the mass gain $\\dmm$ carries the same sense of specific angular momentum, the hole always spins rapidly up to Kerr parameters $a$ close to unity (e.g. Volonteri et al., 2005; Volonteri \\& Rees, 2005). The resulting high radiative efficiency of accretion means that the Eddington limit severely restricts the rate of black hole mass growth, creating difficulties in understanding observations of high masses at early cosmic times. However it is sigmificant that $\\dmm$ is only a very small part of the total mass involved in a major galaxy merger, in line with the observation that the black hole mass in nearby galaxies is typically only $\\sim 10^{-3}$ of the galaxy bulge mass. Moreover, current cosmological simulations are entirely unable to resolve the hydrodynamics of matter accretion on to the central black hole. The mass inflow time through an accretion disc of 1 pc radius is already $\\ga 10^9$~yr (Shlosman, Begelman \\& Frank, 1990), a factor $\\ga 10$ greater than the timescale on which rapid hole growth occurs. The best cosmological simulations resolve only lengthscales at least 100 times larger. Inevitably the simulations have to resort to sub--resolution recipes in trying to model the accretion process. For example it is usual to assume accretion at the Bondi rate from some large radius, whereas in reality the infalling gas must possess angular momentum which will complicate things. In this picture the very small mass $\\dmm$ which accretes on to the central black hole must have almost zero angular momentum about it. Moreover, since the hole is growing its mass as rapidly as possible, it must be close to its Eddington limit and thus feeding back energy and momentum into its surroundings. Indeed simple models of the resulting momentum--driven feedback on to the host galaxy correctly reproduce the $M - \\sigma$ relation between black hole mass and velocity dispersion without any free parameter (King, 2003; 2005, Murray et al., 2005). Along with this, there must be copious star formation from material close to the hole, i.e. at distances $\\sim 0.01 - 0.1$~pc (see below) where the angular momentum neglected in estimating the Bondi accretion rate becomes significant in supporting the captured gas against gravity. This star formation also injects energy and momentum into the gas near the hole. Hence although it is very difficult to predict the gas flow near the black hole it seems more likely to be chaotic than well--ordered, and rather inefficient in growing the hole compared with forming stars, thus keeping the latter a fairly constant fraction $\\sim 10^{-3}$ of the stellar bulge mass. An obvious reason for the relative inefficiency of hole growth versus star formation is that an accretion disc becomes self--gravitating at radii $R$ where its mass exceeds $M_{\\rm sg} \\sim (H/R)M$, where $H$ is the disc scaleheight and $M$ the central (here black hole) mass (e.g. Pringle, 1981, Frank et al. 2002). For AGN this means that discs must be self--gravitating outside some radius $R_{\\rm sg} \\sim 0.01 - 0.1$~pc (Shlosman et al., 1990; Collin--Souffrin \\& Dumont, 1990; Hur\\'e et al., 1994). Cooling times in these regions are so short that self--gravity is likely to result in star formation (Shlosman \\& Begelman, 1989; Collin \\& Zahn, 1999). As discussed by King \\& Pringle, 2007 (hereafter KP07) we expect that most of the gas initially at radii $R > R_{\\rm sg}$ is either turned into stars, or expelled by those stars which do form, on a rapid (almost dynamical) timescale, while the gas which is initially at radii $R < R_{\\rm sg}$ forms a standard accretion disc, which slowly drains on to the black hole and powers the AGN. KP07 show that for local, low--luminosity AGN, fuelling by {\\it well separated} episodes of this type explains observational features such as the luminosity function for moderate--mass black holes, and the presence and location of a ring of young stars observed about the Galactic Centre. This paper deals not with such low--luminosity AGN but instead with high--luminosity events causing major black--hole mass growth. In these cases our discussion suggests a picture in which, within an event, the flow on to the hole is episodic, via a rapid succession of accretion discs limited by self--gravity (as in the well--separated episodes in low--luminosity AGN), and thus with masses $\\dme \\sim \\msg$. The number of such episodes is evidently $\\sim \\dmm/\\msg$. Given that the flow is constantly stirred up by energy and momentum input, and that we are concerned only with the very small fraction of the gas with almost zero angular momentum, it is reasonable to expect the disc orientations to be essentially random. We thus arrive at a picture of AGN accretion close to that suggested by Sanders (1984) (see also Heckman et al., 2004, Greene \\& Ho, 2007, and KP07). This picture reproduces several characteristic features of nearby AGN. In particular the random orientations of the accretion episodes mean that the radio jets show no correlation with the grand--design structure of the host, as observed for low--redshift AGN (Kinney et al., 2000), and inferred at higher redshift by Sajina et al (2007). Moreover the picture implies a luminosity function in broad agreement with that observed for moderate mass central black holes (KP07). Finally this picture provides a plausible explanation for the ring(s) of young stars seen around the black hole in the centre of our own Galaxy (Genzel et al., 2003; Lu et al., 2006). Since observational properties such as the luminosity function of AGN and quasars are sensitive only to the total mass increase $\\Delta M_{\\rm merger}$, and not to its internal angular momentum budget, our picture predicts the same values for these observables as found in previous studies (e.g. Wyithe \\& Loeb, 2003) and is therefore in broad agreement with expected galaxy merger rates. Clearly the proper hydrodynamic treatment of the kind of complex event we envisage must await advances in computing power. However we can gain some insight by simple arguments. ", "conclusions": "We have argued that accretion on to supermassive black holes in AGN occurs through events with total mass $\\dmm \\sim M$, consisting of long sequences of randomly oriented disc accretion episodes with masses $\\dme \\sim 10^{-3}M$. The mass of these disc episodes is strongly constrained by the generic tendency of such discs to become self--gravitating and form stars outside quite small radii $\\sim 0.01 - 0.1$~pc. The self--gravity constraint also limits the disc angular momentum. This is always less than the maximum that the hole's mass would allow it to have. A hole with significant spin therefore stably co-- and counter--aligns initially prograde and retrograde discs respectively, producing a net spindown until its angular momentum is reduced to a relatively modest value, with fluctuations $\\Delta a \\sim \\pm 0.2$ about the average $\\bar a$ given by (\\ref{av}). We note from (\\ref{av}) that $\\bar a$ itself decreases slowly as the black hole mass decreases. We can compare these results with those for coalescences of SMBH as a consequence of the mergers of their host galaxies. Hughes and Blandford (2003) show that the average secular evolution of black hole spin under coalescences is \\begin{equation} a = a_0\\biggl({M_0\\over M}\\biggr)^{2.4} \\end{equation} where $a_0$ is the original spin parameter of a hole within initial mass $M_0$. Thus coalescences doubling the mass of a hole on average reduce $a$ by a factor $\\sim 5.3$. This is an even faster decline of spin than given by assuming fixed angular momentum but growing mass, which would produce $a \\propto M^{-2}$, reflecting the fact that retrograde coalescences are more effective in reducing spin than prograde ones are in growing it. Evidently coalescences can reduce $a$ quite quickly to zero. However we have seen that accretion very rapidly pushes $a$ back towards the mean value $\\bar a$ (cf eqns \\ref{equil}, \\ref{av}) from any starting point. Hence on average coalescences have little net effect on the mean value of $a$. We conclude that repeated accretion episodes and coalescences on average drive the black holes in AGN towards fairly low values of $a$, with statistical fluctuations around them of order $\\Delta a = \\pm 0.2$. Given this result, we should consider the observational evidence constraining values of $a$ for SMBH. Much of this is derived from the Soltan (1982) argument that the totality of background light in the low--redshift Universe is compatible with black--hole growth with radiative efficiency $\\epsilon \\simeq 0.1$ (e.g. Wang et al., 2006; Treister \\& Urry, 2006; Hopkins, Richard \\& Hernquist, 2007), which formally requires $a \\simeq 0.67$. However $\\epsilon$ is such a slow function of $a$ except near $a=1$ (see Figure \\ref{epsilon}) that this conclusion is vulnerable to the systematic effects inherent in trying to estimate the efficiency from observation. Evidently we have to find $\\epsilon \\simeq 0.1$ from observation to an accuracy of a factor $\\la 2$ in order to exclude {\\it any} value of $a$ other than those very near unity. Since retrograde and prograde accretion appear equally probable, the effective efficiency for the Soltan argument is actually the symmetrized curve \\begin{equation} \\epsilon_{\\pm} = {1\\over 2}[\\epsilon(a)+\\epsilon(-a)] \\label{epspm} \\end{equation} in Figure (\\ref{epsilon}), whose vertical range is now compressed to $0.057 - 0.230$. In this case only a very accurate estimate of $\\epsilon$ can give real information about $a$ at all. \\begin{figure} \\centerline{\\epsfxsize9cm \\epsfbox{epsilonbw.eps}} \\caption{Accretion efficiency $\\epsilon$ versus spin parameter $a$ (solid curve). Negative values of $a$ denote retrograde accretion. The dashed curve shows the effective $\\epsilon=\\epsilon_{\\pm}$ given by assuming that retrograde and prograde accretion are equally probable. This is the efficiency relevant to the Soltan argument.} \\label{epsilon} \\end{figure} Streblyanska et al. (2005) suggest a different method of estimating the typical spin $a$ for SMBH. They find that the composite X--ray spectrum of a sample of AGN shows evidence for a broad iron line. They fit lines produced with assumed emissivity and line centre energy from both non--rotating and rotating black holes, and argue that the outer radius of the emission region for the non--rotating hole is unreasonably small in the case of the Type I AGN in their sample, although the statistical fit is marginally preferred over a rotating hole. The best--fitting rotating hole for this group has an inner emission radius of $3GM/c^2$, which requires $a > 0.78$. Since holes with such values of $a$ would automatically have higher efficiencies and thus presumably stronger iron lines, it is difficult to know what fraction of the SMBH in the sample actually have such fairly large spin parameters (note that, as discussed below, our arguments only refer to the statistical mean $\\bar a$, and individual SMBH can differ significantly from this). We remark as a general caveat that the search for any effect depending on $a$, or the interpretation of a particular observational effect as being due to $a$, is likely to produce a bias towards $a \\sim 1$. This effect is particularly marked because of the very high efficiency of prograde accretion for $a \\simeq 1$ (see Fig \\ref{epsilon}). We note that Brenneman \\& Reynolds (2006) estimate $a = 0.989^{+0.009}_{-0.002}$ in the specific case of the broad iron line in MCG-06-30-15, remarkably close to the maximum allowed value. Of course since our treatment is statistical it does not rule out larger spin values in individual cases. The most important way of producing large values of $a$ in individual cases is in coalescences of two SMBH. Hughes \\& Blandford (2003) show that this can produce significant $a$ for prograde coalescences of two black holes with similar mass. This suggests that giant ellipticals offer a promising site for significant values of $a$. Another conceivable way of producing exceptions to our general conclusions is that we have assumed accretion from thin discs, in which cooling is efficient, so that the total disc mass is limited to $\\sim H/R$ times that of the hole. In some situations it may happen that cooling is inefficient and our conclusions do not hold. More work is needed here, although such cases are likely to be rare. Finally one might argue that in some cases an accretion event with mass $\\gg M_{\\rm sg}$ may contrive to retain some memory of an overall angular momentum, despite the fact that it must have a near--radial orbit within the host galaxy in order to accrete at all (cf. Kendall et al., 2003: recall that accretion within a Hubble time requires disc radii $\\la 1$~pc). In this case individual episodes no longer have uncorrelated angular momenta as we assumed. This is effectively what is assumed by Volonteri \\& Rees (2005), and Volonteri et al (2007) for {\\it all} episodes." }, "0801/0801.4177_arXiv.txt": { "abstract": "The Newtonian solid-mechanical theory of non-compressional spheroidal and torsional nodeless elastic vibrations in the homogenous crust model of a quaking neutron star is developed and applied to the modal classification of the quasi-periodic oscillations (QPOs) of X-ray luminosity in the aftermath of giant flares in SGR 1806-20 and SGR 1900+14. Particular attention is given to the low-frequency QPOs in the data for SGR 1806-20 whose physical origin has been called into question. Our calculations suggest that unspecified QPOs are due to nodeless dipole torsional and dipole spheroidal elastic shear vibrations. ", "introduction": "This study was undertaken as a part of current extensive theoretical investigations of asteroseismology of a neutron star\\cite{P-05,GSA-06,SA-07,Lee-07,Lev-07,Bast-07} (see also references therein) that have been boosted by recent discovery of quasi-periodic oscillations (QPOs) of X-ray luminosity in the decaying flares of two magnetars\\cite{Isr-05,SW-06}, SGR 1806-20 and SGR 1900+14, with concomitant suggestion to interpret this variability as caused by quake induced differentially rotational, torsional, oscillations. Following this suggestion the focus of most theoretical works is on computing the frequency spectra of odd-parity torsional mode of shear vibrations and less attention is paid to the even parity spheroidal elastic mode. However, from the viewpoint of modern global seismology\\cite{LW-95,AR-03}, the spheroidal vibrational mode in a solid star and planet has the same physical significance as the toroidal one in the sense that these two fundamental modes owe their existence to one and the same restoring force\\cite{McD-88,Bast-99,Bast-IJ-07}. In this light there is a possibility that, by not considering both these modes on an equal footing, we may miss discovering certain essential novelties which are consequences of solid mechanical laws governing seismic vibrations of superdense matter of neutron stars. Adhering to this attitude and continuing our current investigations\\cite{Bast-IJ-07}, we derive here spectral equations for the frequency of both spheroidal and torsional elastic nodeless vibrations in the solid crust of quaking neutron star and examine what conclusions can be drawn regarding low-frequency QPOs whose physical nature still remain unclear. ", "conclusions": "The obtained spectral formulae for the frequency of nodeless elastic vibrations trapped in the finite-depth seismogenic layer may be of some interest in its own right from the viewpoint of general theoretical seismology\\cite{LW-95} in the sense that they can be utilized in the study of seismic vibrations of more wide class of solid celestial objects such as Earth-like planets. One of the remarkable feature of considered model is that the dipole nodeless overtones possess properties of Goldstone soft modes, that is, the dipole overtones emerge if and on only if the elastic vibrations turn out to be locked in the peripheral layer of finite thickness. It is shown that obtained spectral equations are consistent with the existence treatment of low-frequency QPOs in the X-ray luminosity of flares SGR 1900+14 and SGR 1806-20 as caused by quake-induced torsional nodeless vibrations. What is newly disclosed here is that previously non-identified low-frequency QPOs in data for SGR 1806-20 can be attributed to nodeless dipole torsional and spheroidal vibrations, namely, $\\nu(_0t_1)=18\\,\\mbox{Hz}$ and $\\nu(_0s_1)=26\\,\\mbox{Hz}$." }, "0801/0801.4388_arXiv.txt": { "abstract": "Studies of nearby elliptical and S0 galaxies reveal that roughly half of the low mass X-ray binaries (LMXBs), which are luminous tracers of accreting neutron star or black hole systems, are in clusters. There is a surprising tendency of LMXBs to be preferentially associated with metal-rich globular clusters (GCs), with metal-rich GCs hosting three times as many LMXBs as metal-poor ones. There is no convincing evidence of a correlation with GC age so far. In some galaxies the LMXB formation rate varies with GC color even within the metal-rich peak of the typical bimodal cluster metallicity distribution. This provides some of the strongest evidence to date that there are metallicity variations within the metal-rich GC peak, as is expected in hierarchical galaxy formation scenarios. We also note that apparent correlations between the interaction rates in GCs and LMXB frequency may not be reliable because of the uncertainties in some GC parameters. We argue in fact that there are considerable uncertainties in the integrated properties of even the Milky Way clusters that are often overlooked. ", "introduction": "Chandra X-ray images of nearby ellipticals and S0s reveal large numbers of point sources, confirming a long-standing suggestion that the hard X-ray emission in many of these galaxies comes predominantly from X-ray binaries. In the old stellar populations of these galaxies the bright, L$_X$$\\geq$10$^{37}$ erg s$^{-1}$ sources detected in typical Chandra observations are low mass X-ray binaries, binary systems comprising a neutron star or black hole accreting via Roche lobe overflow from a low mass companion. An interesting characteristic of LMXBs is that they are disproportionately abundant in globular clusters. In the Milky Way (MW) the probability of finding a LMXB among the stars in the globular clusters is a few hundred times larger than for field stars. This has long been attributed to efficient formation of LMXBs in GCs due to dynamical interactions in the core. Early type galaxies are ideal for studies of the LMXB-GC link as they are particularly abundant in clusters. While there are $\\approx$150 globular clusters in the Milky Way, bright elliptical galaxies typically host several thousand GCs. Globular clusters are among the simplest stellar systems known; Thus the stellar population in each of these compact, $\\sim$10$^5$ M$_\\odot$, luminous systems has well defined characteristic properties such as age and metallicity. The identification of LMXBs with GCs provides a unique opportunity to probe the effects of these parameters on LMXB formation and evolution. \\begin{figure}[!ht] \\begin{center} \\centerline{\\psfig{figure=kundu_a_fig1.ps,width=4.5in}} \\caption{ Top: The V-I colors of GCs vs. distance from the center of NGC 4472 and GC color distribution. LMXB-GC matches are indicated by large filled circles/bins. The colors of these clusters primarily trace the underlying (bimodal) metallicity distribution. Bottom: V magnitude of globular clusters vs. half light radius and the globular cluster luminosity function. LMXBs are preferentially located in the high mass and metal-rich GCs. There is a weak anti-correlation with GC half-light radius, but no obvious correlation with galactocentric distance. Each of these trends has been confirmed in other galaxies. } \\end{center} \\end{figure} ", "conclusions": "" }, "0801/0801.2084_arXiv.txt": { "abstract": "We performed a photometric multicolor survey of the core of the Canis Major over-density at $l \\approx 244^{o}$, $b\\approx -8.0^{o}$, reaching V $\\sim$ 22 and covering $0^{o}.3 \\times 1^{o}.0$. The main aim is to unravel the complex mixture of stellar populations toward this Galactic direction, where in the recent past important signatures of an accretion event have been claimed to be detected. While our previous investigations were based on disjointed pointings aimed at revealing the large scale structure of the third Galactic Quadrant, we now focus on a complete coverage of a smaller field centered on the Canis Major over-density. A large wave-length baseline, in the $UBVRI$ bands, allows us to build up a suite of colour colour and colour magnitude diagrams, providing a much better diagnostic tool to disentangle the stellar populations of the region. In fact, the simple use of one colour magnitude diagram, widely employed in all the previous studies defending the existence of the Canis Major galaxy, does not allow one to separate the effects of the different parameters (reddening, age, metallicity, and distance) involved in the interpretation of data, forcing to rely on heavy modeling. In agreement with our previous studies in the same general region of the Milky Way, we recognize a young stellar population compatible with the expected structure and extension of the Local (Orion) and Outer (Norma-Cygnus) spiral arms in the Third Galactic Quadrant. Moreover we interpret the conspicuous intermediate-age metal poor population as belonging to the Galactic thick disk, distorted by the effect of strong disk warping at this latitude, and to the Galactic halo. ", "introduction": "In the last years, we have used photometric observations of young open cluster fields to probe the spiral structure in the third Galactic Quadrant (TGQ, $180^{o} \\leq l \\leq 270^{o}$; Carraro et al. 2005a, Moitinho et al. 2006, V\\'azquez et al. 2008), motivated by the very poor knowledge of this portion of the Galaxy's periphery. Interestingly, important low latitude accretion phenomena have been recently claimed to be ongoing in this part of the Galaxy, such as the Canis Major over-density (CMa, Bellazzini et al. 2004), and the Monoceros Ring (MRi, Newberg et al. 2002). Clearly, a detailed description of the structure and stellar populations of the Galactic disc (thin plus thick) is mandatory to discriminate between Galactic and extragalactic material. The TGQ is a special region of the Milky Way's outskirts, characterized by significant absorption windows as the Puppis (l $\\sim 243^{o}$) window (Fitzgerald 1968, Moffat et al. 1979, Janes 1991, Moitinho 2001), which allows one to detect very distant star clusters (Baume et al. 2006). Besides, and interestingly, young star clusters are found at low Galactic latitudes, underlining the fact that the young Galactic disk is significantly warped in these directions (May et al. 1997, Momany et al. 2004, Moitinho et al. 2006, Momany et al. 2006, L\\'opez-Corredoira et al. 2007). In V\\'azquez et al (2008), by combining optical and CO observations, we have provided a fresh and very detailed picture of the spiral structure in the TGQ, showing that this region is characterized by a complicated spiral pattern. The outer (Norma-Cygnus) arm is found to be a grand design spiral feature defined by young stars, whereas the region closer to the Sun (d$_{\\odot}$ less than 9 kpc) is dominated by a conspicuous inter-arm structure, at l $\\sim$ 245$^o$, the Local spiral arm. In this region, Perseus is apparently defined by gas and dust, and does not appear to be traced by an evident optical young stellar population, similarly to what can be found in other galaxies such as M~74 (V\\'azquez et al. 2008). The analysis carried out on a substantial fraction of the stellar fields we observed revealed a complicated mixture of young and old populations. Although centered on catalogued star clusters (Dias et al. 2002), a few colour magnitude diagrams (CMD) do not reveal star clusters but, and more interestingly, show hints of a young, diffuse, and distant stellar populations, which has become recently referred to as {\\it Blue Plume} ( Bellazzini et al. 2004, Dinescu et al. 2005, Mart\\'inez-Delgado et al. 2005,Carraro et al. 2005 ). Since the disk is warped and flared in these directions (Momany et al. 2006), the lines of sight are expected to cross both the thin and thick disk population in front of a particular target, in a way that the analysis of the CMD becomes very challenging (see for instance the analysis of the field toward the star cluster Auner~1, Carraro et al. 2007). In this paper we present a photometric analysis in the $UBVRI$ filters of the stellar populations in 3 wide field pointings toward the CMa over-density. Sect.~2 describes the observation and data reduction strategies. In Sect.~3 we discuss various colour combination CMDs, while Sects.~4 and 5 are dedicated to illustrate and analyze the TCD as a function of magnitude. Finally, Sect.~6 summarizes our findings. \\begin{table*} \\centering \\begin{tabular}{cccccccccccc} \\hline\\hline Designation & $\\alpha (2000.0)$ & $\\delta(2000.0)$ & l & b & U & B & V & R & I & Airmass & Seeing\\\\ & & & [deg] & [deg] & secs & secs & secs & secs & secs & & arcsec\\\\ \\hline Field 1&07:22:51&-30:59:20&244.00&-07.50&20,180,1800&10,150,1500&5,60,900&5,60,900&5,60,900&1.00-1.30&0.83-1.02\\\\ Field 2&07:20:46&-31:09:36&244.00&-08.00&20,180,1800&10,150,1500&5,60,900&5,60,900&5,60,900&1.00-1.30&0.90-1.12\\\\ Field 33&07:20:21&-31:15:43&244.00&-08.10&20,180,1800&10,150,1500&5,60,900&5,60,900&5,60,900&1.20-1.80&1.28-2.11\\\\ \\hline\\hline \\end{tabular} \\caption{ List of pointings discussed in this paper. For each pointing, equatorial and galactic coordinates are reported together with the set of filters used, and the range of exposure time, air-mass and typical seeing. The fields are shown in Fig.~1 to 3} \\end{table*} \\section[]{Observations and Data Reduction} \\subsection{Observational details} $UBVRI$ images of 3 overlapping fields (see Figs.~1 to 3) in the Third Quadrant of the Milky Way toward the CMa over-density were obtained at the Cerro Tololo Inter-American Observatory 1.0m telescope, which is operated by the SMARTS\\footnote{http://www.astro.yale.edu/smarts/} consortium. The telescope is equipped with a new 4k$\\times$4k CCD camera having a pixel scale of 0$^{\\prime\\prime}$.289/pixel which allows to cover a field of $20^{\\prime} \\times 20^{\\prime}$ on the sky. Observations were carried out on the nights of November 28 and December 3, 2005. The two nights were part of a 6 night run. In the first night we observed the fields $\\#1$ and $\\#2$ (see Table~1) while field $\\#3$ was observed in the last night of the run. The CCD is read out through 4 amplifiers, each one with slightly different bias levels and gains. Pre-processing was done using the procedure developed by Philip Massey\\footnote{http://www.lowell.edu/users/massey/obins/y4kcamred.html}. Briefly, the procedure trims and corrects the images for bias, flat-field, and bad pixels, preparing them from photometric extraction. A series of skyflats was employed in all the filters. \\begin{figure} \\includegraphics[width=\\columnwidth]{MN1806fig4.eps} \\caption{Photometric solution in $UBVRI$ for standard stars. See Table~2 for details. $\\sigma$, on the right, indicates the {\\it rms} of the fit.} \\end{figure} \\begin{table} \\caption{Calibration coefficients.} \\begin{tabular}{rrr} \\hline \\multicolumn {3}{c}{$u_1 = +3.292 \\pm 0.005$, $u_2 = +0.026 \\pm 0.006$, $u_3 = +0.49$} \\\\ \\multicolumn {3}{c}{$b_1 = +2.187 \\pm 0.004$, $b_2 = -0.164 \\pm 0.005$, $b_3 = +0.25$} \\\\ \\multicolumn {3}{c}{$v_1 = +1.930 \\pm 0.003$, $v_2 = +0.010 \\pm 0.003$, $v_3 = +0.16$}\\\\ \\multicolumn {3}{c}{$r_1 = +1.936 \\pm 0.004$, $r_2 = -0.012 \\pm 0.009$, $r_3 = +0.09$}\\\\ \\multicolumn {3}{c}{$i_1 = +2.786 \\pm 0.004$, $i_2 = +0.015 \\pm 0.004$, $i_3 = +0.08$} \\\\ \\hline \\end{tabular} \\end{table} \\subsection{Standard Stars} Three Landolt (1992) areas (TPhoenix, Rubin~149, and PG~0231+006) were observed several times each night to tie instrumental magnitudes to the standard system. All nights, except the last one, were stable and photometric with seeing between 0.8 and 1.2 arcsec. The last night was non-photometric with bad seeing conditions (see Table~1). Photometry from this last night was tied to the other nights through the comparison of stars in common.\\\\ Since the photometric solutions were identical, all the standard star measurements were used together in obtaining a single photometric solution for the entire run. This resulted in calibration coefficients derived using about 200 standard stars. Photometric solutions have been calculated following Patat \\& Carraro (2001). Fig.~4 shows the run of magnitude differences (standard versus instrumental) for the whole standard set. Notice that the colour baseline is sufficiently broad. On the right, the {\\it rms} of the fit is shown for each colour. The calibration equations read: \\begin{center} \\begin{tabular}{lc} $u = U + u_1 + u_2 (U-B) + u_3 X$ & (1) \\\\ $b = B + b_1 + b_2 (B-V) + b_3 X$ & (2) \\\\ $v = V + v_1 + v_2 (B-V) + v_3 X$ & (3) \\\\ $r = R + r_1 + v_2 (V-R) + r_3 X$ & (4) \\\\ $i = I + i_1 + i_2 (V-I) + i_3 X$ & (5), \\\\ \\end{tabular} \\end{center} \\noindent where $UBVRI$ are standard magnitudes, $ubvri$ are the instrumental magnitudes, $X$ is the airmass, and the derived coefficients are presented in Table~2. We adopted the extinction coefficients typical of the site (Carraro et al. 2005b). \\subsection{Photometry extraction} The covered areas are shown in Figs.~1 to 3. Data have been reduced using IRAF\\footnote{IRAF is distributed by NOAO, which is operated by AURA under cooperative agreement with the NSF.} packages CCDRED and DAOPHOT. Photometry was done employing the point spread function (PSF) fitting method (Stetson 1987). Particular care has been put in defining the PSF model. A variable PSF was adopted due to PSF variations across the CCD. In general, up to 40 bright stars have been selected for defining the PSF model. Aperture corrections were estimated from samples of bright PSF stars (typically 15), and then applied to all the stars. The corrections amounted to 0.250-0.315, 0.280-0.300, 0.200-0.280, 0.190-0.270, and 0.210-0.280 mag for the $UBVRI$ filters, respectively, over the entire run. Photometric completeness was estimated following Baume et al. (2006) and was determined to be higher than 50$\\%$ at V $\\sim$ 20. mag. ", "conclusions": "We have presented a photometric analysis in the $UBVRI$ filters of 3 wide field pointings close to the center of the Canis Major over-density. The goal was to study the stellar populations in this region of the Milky Way, where a putative dwarf galaxy in the act of being cannibalized by the Milky Way, is claimed to exist. The analysis presented in this paper followed a different strategy from previous investigations of the CMa over-density. Instead of studying very large fields in two filters, we have concentrated on a smaller area, but observed in several filters. This approach, frequently employed in the study of star clusters, allowed us to construct several CMDs and the classical (B-V) vs (U-B) TCD, which together constitute a very powerful tool for detecting young stellar populations. As in our previous studies of stellar fields in the TGQ we found evidence of a diffuse young stellar population, as expected from the presence of the Local and Outer Galactic spiral arms (Carraro et al. 2005, Moitinho et al. 2006, V\\'azquez et al. 2008). Once again, no indication has been found of an ongoing accretion event in this direction of the Galactic disk. In addition, the estimated ranges of distance, age and metallicity of the older metal poor population are consistent with those of thick disk stars at different distances from the Sun. These findings, together with the results of previous papers by us and other authors (Momany et al 2004, 2007), significantly weaken the proposed scenario of a dwarf galaxy in CMa being cannibalized by the Milky Way. Instead, all the observational evidence fits our current knowledge of the Galactic disk. The TGQ is indeed a complicated region due to the warp and the existence of the Local Arm. Only the detailed multicolour analysis we have been conducting in the last few years could provide a clear picture of the structure of the outer disk in the TGQ. \\begin{figure} \\includegraphics[width=\\columnwidth]{MN1806fig12.eps} \\caption{V vs V-I CMD of the CMa over-density. Two isochrones (dashed lines, red when printed in color) have been superimposed to guide the eye, illustrate the position of the TO, and provide a rough estimate of the mean age of the population.} \\end{figure}" }, "0801/0801.0424_arXiv.txt": { "abstract": "In the local Universe, high-power radio galaxies live in lower density environments than low-luminosity radio galaxies. If this trend continues to higher redshifts, powerful radio galaxies would serve as efficient probes of moderate redshift groups and poor clusters. Photometric studies of radio galaxies at $0.3 \\lesssim z \\lesssim 0.5$ suggest that the radio luminosity-environment correlation disappears at moderate redshifts, though this could be the result of foreground/background contamination affecting the photometric measures of environment. We have obtained multi-object spectroscopy of in the fields of 14 lower luminosity ($L_{\\rm{1.4~GHz}} < 4\\times10^{24}~\\rm{W~Hz^{-1}}$) and higher luminosity ($L_{\\rm{1.4~GHz}} > 1.2\\times10^{25}~\\rm{W~Hz^{-1}}$) radio galaxies at $z \\approx 0.3$ to spectroscopically investigate the link between the environment and the radio luminosity of radio galaxies at moderate redshifts. Our results support the photometric analyses; there does not appear to be a correlation between the luminosity of a radio galaxy and its environment at moderate redshifts. Hence, radio galaxies are not efficient signposts for group environments at moderate redshifts. ", "introduction": "Photometric studies of radio galaxies have shown that Fanaroff-Riley type I \\citep[FR I;][]{fr} (low luminosity and edge-dimmed radio lobes) and FR II (high luminosity and edge-brightened lobes) radio galaxies exist in different environments in the local Universe; FR I galaxies reside in rich clusters while FR II galaxies tend to inhabit poor clusters or rich groups \\citep[e.g.,][]{heckman,prestage,lilly}. Studies of higher redshift radio galaxies indicate that this difference disappears at moderate redshifts ($z \\sim 0.5$), where both FR types are associated with cluster-like environments \\citep[e.g.,][]{hill,bahcall,allington}. In contrast, \\citet{zirbel} suggests that, although the environments of both types of galaxies are more dense at moderate redshifts than low redshifts, the environments of FR II galaxies remain less dense than FR I sources out to $z \\sim 0.5$. Only the low redshift trend of FR I environments being more dense than the environments of FR II galaxies has been confirmed with X-ray observations \\citep{miller99} and optical spectroscopy \\citep{miller}. It is quite difficult to detect and characterize group-like over-densities at moderate redshifts without spectroscopy. Photometric properties are difficult to use because groups tend not to have substantial bright early-type populations \\citep[e.g.,][]{wilman} that form a red sequence, and foreground and background interlopers are common. Therefore, field spectroscopy is necessary to determine whether FR II galaxies may be useful signposts to identify moderate redshift galaxy groups. These galaxy groups and poor clusters are excellent laboratories for studying galaxy evolution. Over half of all galaxies in the local Universe are members of groups \\citep{huchra,tully}, and these groups have moderate densities and low velocity dispersions that make them the preferred environments for dynamic galaxy evolution \\citep[e.g.,][]{zabludoff,carlberg,aarseth,barnes} compared to clusters or the field \\citep[e.g.,][]{aceves,lin,conselice,patton}. Though cluster cores also promote some galaxy processing and evolution, the high velocity dispersions associated with these cores decrease the effective interaction cross-section between galaxies and limit the evolutionary effects of galaxy-galaxy interactions \\citep[e.g.,][]{bower}. Therefore galaxy groups provide excellent conditions for studying interactions that affect the star formation rate of galaxies and AGN ignition \\citep[e.g.,][]{mihos,iono,bahcallj,best}. The most significant obstacle to studying galaxy evolution in group environments is identifying non-local groups. The Sloan Digital Sky Survey \\cite[SDSS;][]{york} and the 2 Degree Field Galaxy Redshift Survey \\cite[2dFGRS;][]{colless} have extended local group samples \\citep[e.g.,][]{ramella,zabludoff} to redshifts $z \\sim 0.2$ \\citep[e.g.,][]{weinmann,collister,padilla}. Blind redshift surveys at higher redshifts have successfully identified galaxy groups to $z \\sim 0.7$ \\citep[e.g.,][]{wilman,gerke}, but the substantial amount of telescope time invested in these surveys makes it desirable to find alternative methods of identifying moderate redshift groups. \\citet{mulchaey} use deep X-ray observations to select groups at moderate redshifts, though this method may preferentially select the most massive groups. Gravitational lenses have also been used to identify moderate redshift groups \\citep{momcheva,williams,augera,augerb,fassnacht}, though the scarcity of lenses limits the utility of this method. Radio galaxies have been used extensively to find high redshift clusters \\citep[e.g.,][]{deltorn,blantona,blantonb} and protoclusters \\citep[e.g.,][]{miley,venemans}, but the FR I/FR II environment dichotomy suggests that they may also be useful for identifying less rich galaxy groups \\citep[e.g.,][]{allington}. In this paper we present spectroscopic observations of the fields of a sample of 14 moderate redshift radio galaxies. We investigate the relationship between the radio luminosity and spectroscopic group properties, and we analyze the effectiveness of using high luminosity radio galaxies to preferentially locate moderate redshift groups. Throughout this paper we assume a radio spectral index of $\\alpha = 0.5$, where $S_{\\nu} \\propto \\nu^{-\\alpha}$, and we use a $\\Lambda$CDM cosmology with $\\Omega_{\\Lambda} = 0.73$ and $\\Omega_M = 0.27$. ", "conclusions": "\\citet{geach} have examined the environments of a set of four radio galaxies at redshifts $z \\sim 0.35$ and $z \\approx 0.65$ with spectroscopy and X-ray observations. This sample includes three galaxies that we would classify as low-power and one higher-power radio galaxy, and the sample spans a range of environments. We note, however, that none of the galaxies in our sample or the \\citet{geach} sample are truly isolated; all have at least two spectroscopic companions. This could be due to selection effects. \\citet{geach} investigated radio galaxies that appeared to have a red sequence, while our sample required the galaxy to be in the SDSS spectroscopic sample. The SDSS preferentially observes more luminous galaxies at higher redshifts, and these galaxies tend to exist in over-dense environments. However, photometric surveys of purely radio-selected galaxies have also suggested that nearly all radio-galaxies reside in groups \\citep[e.g.,][]{allington}, and we therefore do not expect this bias to be significant. Our results spectroscopically support the results from photometric surveys of radio galaxies; low-luminosity and high-luminosity sources tend to exist in the same environments at moderate redshifts. The dichotomy in environments at low redshifts is likely due to the density of the intergalactic medium (IGM); high luminosity galaxies tend to have large extended lobes that do not form if the IGM is too dense \\citep[e.g.,][]{prestage}. The absence of this dichotomy at higher redshift is probably results from cosmological evolution of the density of the IGM \\citep[e.g.,][]{ettori}. There is a substantial amount of scatter in the properties of the radio galaxy environments, though the typical radio galaxy is in a group with approximately 12 members and a velocity dispersion of $\\sim 500~\\rm{km~s^{-1}}$. The systems 222821.192$+$011412.29 and 013352.730$+$011343.63 are found to have only 3 companions in addition to the radio galaxy. These systems may not belong to bound groups but rather may be embedded in filaments; the large velocity spread for the low-luminosity system 222821.192$+$011412.29 further supports this conclusion. We therefore conclude that at moderate redshifts, radio galaxies reside in all types of environments, from filaments to clusters, without regard to the radio luminosity. \\citet{belsole} reach a similar conclusion from a study of the X-ray environments of higher redshift radio galaxies and quasars. Thus, while radio galaxies are shown to trace large-scale-structure, they are not useful for targeting environments with specific properties (e.g., lower velocity dispersion or less rich environments) at higher redshifts." }, "0801/0801.1034_arXiv.txt": { "abstract": "{Using extremely deep (rms $\\sim$3.3$\\mu$Jy/bm) 1.4\\,GHz sub-arcsecond resolution MERLIN$+$VLA radio observations of a 8\\farms5$\\times$8\\farms5 field centred upon the Hubble Deep Field North, in conjunction with {\\it Spitzer} 24\\,$\\mu$m data we present an investigation of the radio-MIR correlation at very low flux densities. By stacking individual sources within these data we are able to extend the MIR-radio correlation to the extremely faint ($\\sim$microJy and even sub-microJy) radio source population. Tentatively we demonstrate a small deviation from the correlation for the faintest MIR sources. We suggest that this small observed change in the gradient of the correlation is the result of a suppression of the MIR emission in faint star-forming galaxies. This deviation potentially has significant implications for using either the MIR or non-thermal radio emission as a star-formation tracer at low luminosities. } \\FullConference{From planets to dark energy: the modern radio universe\\\\ October 1-5 2007\\\\ University of Manchester, Manchester, UK} \\begin{document} ", "introduction": "Radio and infrared emission from galaxies in both the nearby and distant Universe is thought to arise from processes related to star-formation, hence resulting in the correlation between these two observing bands. The infrared emission is produced from dust heated by photons from young stars and the radio emission predominately arises from synchrotron radiation produced by the acceleration of charged particles from supernovae explosions. It has however recently been suggested that at low flux density and luminosities there may be some deviation from the tight well-known radio-IR correlation seen for brighter galaxies \\cite{bell03,boyle07}. \\cite{bell03} argue that while the IR emission from luminous galaxies will trace the majority of the star-formation in these sources, in low luminosity galaxies the IR emission will be less luminous than expected considering the rate of star-formation within the source (i.e. the IR emission will not fully trace the star-formation). In this scenario the reduced efficiency of IR production relative to the source star-formation rate (SFR) would be the result of inherently lower dust opacities in lower luminosity sources and consequently less efficient reprocessing of UV photons from hot young stars into IR emission. The simple consequence of this is that at lower luminosities the near linear radio-IR correlation L$_{\\rm radio}\\propto$L$_{\\rm IR}^{\\!\\gamma}$, with $\\gamma >1$ (e.g. \\cite{cox88,price92}) will be deviated from. {\\it Of course such an assertion is dependent upon the radio emission providing a reliable tracer of star-formation at low luminosities which may be equally invalid.} Recently \\cite{boyle07} have presented a statistical analysis of Australia Telescope Compact Array (ATCA) 20\\,cm observations of the 24\\,\\mum\\ sources within the \\spitzer\\ Wide Field Survey (SWIRE). In this work \\cite{boyle07} have co-added sensitive (rms$\\sim$30\\,$\\mu$Jy) radio data at the locations of several thousand 24\\,\\mum\\ sources. Using this method they have statistically detected the microJy radio counterparts of faint 24\\,\\mum\\ sources. At low flux densities (S$_{\\rm 24\\, \\mu m }=100\\,\\mu$Jy) they confirm the IR-radio correlation but find it to have a lower coefficient (S$_{\\rm 1.4\\,GHz}$\\,=\\,0.039\\,S$_{\\rm 24\\,\\mu m}$) than had previously been reported at higher flux densities. This coefficient is significantly different from results previously derived from detections of individual objects (e.g. \\cite{appleton04}) and is speculated by \\cite{boyle07} to be the result of a change in the slope of the radio-IR correlation at low flux densities. In this work (which is descirbed in more detail in \\cite{beswick06, beswick08}) we have utilised very deep, high resolution 20\\,cm observations of the Hubble Deep Field North and surrounding area made using MERLIN and the VLA \\cite{muxlow05} in combination with publicly available 24\\,\\mum\\ \\spitzer\\ source catalogues from GOODS to study the MIR-Radio correlation for microjansky radio sources. This study extends the flux density limits of the radio-IR correlation by more than an order of magnitude for individual sources and overlaps the flux density regime studied using statistical stacking methods by other authors. \\begin{figure*} \\begin{center}$ \\begin{array}{cc} \\includegraphics[width=8cm]{Beswick-fig3a.ps} & \\includegraphics[width=8cm]{Beswick-fig3b.ps} \\end{array}$ \\caption{Mean (left) and median (right) images of the 1.4\\,GHz radio emission for all sources within the six faintest 24\\,\\mum\\ flux density logarithmic bins plotted in Fig.\\,1 in descending flux density order from top-left to bottom-right. The range of 24\\,\\mum\\ flux density over which each image has been stacked is shown at the top of individual panels. Each image is contoured with levels of $-$2, $-$1.414, $-$1,1, 1.414, 2, 2.828, 4, 5.657, 8, 11.31, 16, 22.63 and 32 times 3$\\times$(3.3/$\\sqrt {\\rm N}$)\\,$\\mu$Jy\\,bm$^{-1}$, where N equals the number of 24\\,\\mum\\ source positions averaged in the map. The peak flux density, lowest plotted contour and number of IR sources which have been averaged over (N) is listed at the bottom of each image panel.} \\label{maps} \\end{center} \\end{figure*} ", "conclusions": "Using one of the deepest high-resolution 1.4\\,GHz observations made to date, in conjunction with deep 24\\,\\mum\\ \\spitzer\\ source catalogues from GOODS, we have investigated the microJy radio counterparts of faint MIR sources. These observations confirm that the microJy radio source population follow the MIR-radio correlation and extend this correlation by several orders of magnitude to very low flux densities and luminosities, and out to moderate redshifts. This extension of the MIR-radio correlation confirms that the majority of these extremely faint radio and 24\\,\\mum\\ sources are predominantly powered by star-formation with little AGN contamination. Statistically stacking the radio emission from many tens of faint 24\\,\\mum\\ sources has been used to characterise the size and nature of the radio emission from very faint IR galaxies well below the nominal radio sensitivity of these data. Using these methods the MIR-radio correlation has been further extended and a tentative deviation in this correlation at very low 24\\,\\mum\\ flux densities has been identified." }, "0801/0801.3177_arXiv.txt": { "abstract": "{ The HST/ACS Survey of Galactic globular clusters (GGCs) is a HST Treasury project aimed at obtaining high precision photometry in a large sample of globular clusters. The homogeneous photometric catalogs that has been obtained from these data by Anderson et al. (2008) represents a golden mine for a lot of astrophysical studies. In this paper we used the catalog to analyse the properties of MS-MS binary systems from a sample of fifty GGCs. We measured the fraction of binaries (divided in different groups), studied their radial distribution and constrained the mass ratio distribution. We investigated possible relations between the fraction of binaries and the main parameters of their host GGCs. We found a significant anti-correlation between the binary fraction in a cluster and its absolute luminosity (mass). ", "introduction": "Knowledge of the binary frequency in Globular Clusters (GCs) is of foundamental importance for a lot of astrophisical studies. Binaries play an important role in the dynamical evolution of a clusters. Interactions with hard binaries pump kinetic energy into the cluster core, slowing the core collaps and, eventually, causing the core to reexpand, if the number of binaries is large enough. In general, binaries are a foundamental ingredient in any dynamical evolution model of a GC. Exotic stellar objects, like Blue Stragglers, cataclismic variables, millisecond pulsars and low mass X ray binaries are believed to represent evolutionary stages of close binary system. The determination of the fraction of binaries plays a foundamental step towards the understanding of the evolution of these peculiar objects. Furthermore, binary stars introduce systematic errors in the determination of the main sequence (MS) fiducial line and move it toward red colors with respect to its correct position. Finally, a correct determination of the mass and luminosity functions requires a correct measure of the fraction of binaries. Up to now, three main techniques have been used to measure the fraction ob binaries in GGCs (Hut et al. 1992). The first one identifies binaries by measuring their radial velocity variation (eg. Latham 1996). This method relies with the detection of each individual binary system but, due to the limits in sensibility of spectroscopy, these studies are possible only for the brightest GGCs stars. The second tecnique is based on the search for photometric variables (eg. Mateo 1996). As well as the previous one, it is able to infer specific properties of each binary system (like the measure of orbital period, mass ratio, orbital inclination). Unfortunatly, it is biased towards binaries with short periods and large orbital inclination. Moreover these tecniques have a low discovery efficency and are very expensive in terms of telescope time because it is necessary to repeat measures in time. A thirth approach, that is based on the analysis of the number of stars located on the red side of the main sequence (MS) ridge line (MSRL) may represent a more efficient method to measure the fraction of binaries in a cluster for several reasons: \\begin{itemize} \\item{availability of a large number (thousands) of stars makes it a statistically roboust method;} \\item{it is cheep in terms of observational time: two filters are enough for detecting binaries and repeated measuremets are not needed.} \\item{it is sensitive to binaries with any orbital period and inclination} \\end{itemize} This approach have been used by other groups ( see Sollima et al. 2007 and references within) to study the population of binaries in GGCs. The relative small number of clusters that have been analized is consequence of the intrinsic difficulties of the method: \\begin{itemize} \\item{high photometric quality is required;} \\item{in some cases, the differential reddening spreads the MS and makes it more difficult to isolate the binary sequence;} \\item{an accurate analysis of photometric errors as well as a correct estimate of field contamination are necessary to disantangle real binaries from bad photometry and field stars. } \\end{itemize} In this paper, we analyse the catalogs obtained by Anderson et al. (2008) from HST ACS/WFC data. We exploited both the homegeneity of this dataset, and the high photometric accuracy of the measures to derive the fraction of binaries in the central regions of fifty GGCs. ", "conclusions": "" }, "0801/0801.4561_arXiv.txt": { "abstract": "Planets embedded in optically thick passive accretion disks are expected to produce perturbations in the density and temperature structure of the disk. We calculate the magnitudes of these perturbations for a range of planet masses and distances. The model predicts the formation of a shadow at the position of the planet paired with a brightening just beyond the shadow. We improve on previous work on the subject by self-consistently calculating the temperature and density structures under the assumption of hydrostatic equilibrium and taking the full three-dimensional shape of the disk into account rather than assuming a plane-parallel disk. While the excursion in temperatures is less than in previous models, the spatial size of the perturbation is larger. We demonstrate that a self-consistent calculation of the density and temperature structure of the disk has a large effect on the disk model. In addition, the temperature structure in the disk is highly sensitive to the angle of incidence of stellar irradition at the surface, so accurately calculating the shape of the disk surface is crucial for modeling the thermal structure of the disk. ", "introduction": "Giant planets forming by core accretion need to have cores of 10-20 M$_{\\earth}$ to be massive enough to accrete a gaseous envelope \\citep{2005Hubickyj_etal}. This predicts that sizeable planet embryos form before circumstellar gas disks dissipate. These disks are typically modeled as passive accretion disks, optically thick and gas-dominated. Their temperature structure is strongly dependent on heating from stellar irradiation \\citep{CG, calvet, vertstruct, dalessio2, dalessio3}. In particular, the disk temperatures depend strongly on the angle of incidence of the stellar irradiation on the disk surface. While a substantial amount of work on numerical hydrodynamic simulations of planets embedded in disks has been carried out, calculating the radiative transfer of stellar irradiation in conjunction with these simulations is prohibitively difficult. Simulations by \\citet{bate} illustrate the effect of hydrodynamics on the disk structure around a small embedded planet, but use an isothermal equation of state. The effects of MHD turbulence have been studied by \\citet{2004MNRAS.350..829P} and \\citet{2007Oishi_etal}, but again, assuming a simple and unrealistic equation of state. \\citet{2006KlahrKley} and \\citet{2006PaardekooperMellema,2008PaardekooperMellema} have made an effort to include radiative transfer as well as hydrodynamics in their calculations, but do not include the effects of stellar irradiaton on their models. The work presented in this paper does not include hydrodynamics, but rather focuses on the effects of stellar irradiation, which is a particularly important source of heating in the photospheres of disks. Since it is the photospheres of optically thick disks that are observed, the effects of stellar irradiation at the surface of disks are an important consideration in predicting observations of protoplanetary disks. A limitation of the hydrodynamic simulations described above is that in order to adequately model the densest regions of the gas they are necessarily limited in spatial range above the midplane. The photosphere and surface of a disk are often several scale heights above the midplane. The models presented here are complementary to detailed hydrodynamic simulations for this reason. We are particularly interested in determining if the growing cores of giant planets produce effects that are observable. If these effects are observed, they would affirm the core accretion model as the paradigm for giant planet formation. While a fully-formed Jupiter-mass planet would produce a larger feature in a disk, it would reveal little about how it formed. Previous work on small planets embedded in optically thick gas disks indicates that sub-Jovian mass planet cores can perturb the disk enough to alter the temperature structure of the disk in the immediate vicinity of the planet, with consequences for further evolution of the planet \\citep{paper1, paper2, HJCmigration}. In \\citet{paper1,paper2} (henceforth Paper I and Paper II, respectively), the planet is predicted to gravitationally compress the disk in the vertical direction, creating a shadow paired with a bright spot, leading to temperature variations. However, those models were limited by being plane-parallel, despite the sizes of the simulation boxes used. In addition, while the density perturbations were calculated under hydrostatic equilibrium, they were not calculated self-consistently with the temperature perturbations. In this paper, we improve upon the model presented in Papers I and II by iteratively calculating the density and temperature structure of the disk for self-consistency and eliminating the assumption of a locally plane parallel model. The paper is organized as follows: in \\S \\ref{model} we describe the basic model and the improvements we have made, in \\S \\ref{stellarheating} we elaborate in detail on the method we use for modeling radiative transfer, in \\S \\ref{results} we describe our results for a different planet masses and distances, in \\S \\ref{discussion} we discuss our results in comparison to previous models, and in \\S \\ref{conclusion} we present our conclusions. ", "conclusions": "This paper presents a model for calculating the density and temperature perturbations imposed on a protoplanetary disk by an embedded protoplanet. The basic radiative transfer model is adopted from Papers I and II, but a number of improvements have been made on that work such as density-temperature self-consistency and eliminating the assumption of a plane-parallel system. We have shown that self-consistently calculating the temperature and density significantly increases the effect of planet perturbations by means of postive feedback, where shadowed regions cool and contract and brightened regions heat and expand. This demonstrates the importance of self-consistency when calculating disk structure with radiative transfer. While it has already been acknowledged that stellar irradiation heating is important in setting overall disk structure \\citep[e.g.][]{CG,thermstab,vertstruct,2004DullemondDominik}, this work shows that it is important for considering local perturbations on the disk. Another important result of this paper is that the temperature structure of the disk is extremely sensitive to the angle of incidence of stellar irradiation at the surface. The precise determination of the surface of the disk is critical to accurately calculating the temperature structure of the disk. In previous work we had assumed that the surface of constant density was a sufficiently good approximatation for the surface, but in this work we show that that overestimated the temperature perturbation to the disk. We have omitted some important physics in this calculation. We do not include heating from accretion onto the planet. We consider the embedded planet to act simply as a gravitational point mass, and focus only on the vertical component of gravity. We do not account for non-linearities such as spiral density waves. All these effects are more adequately addressed using a three-dimensional hydrodynamic simulation of a planet embedded in a disk of gas. However, since simulations of this sort focus on the bulk flow of gas at the midplane, they typically have insufficient resolution above the midplane to accurately calculate the surface of the disk \\citep[e.g.][]{bate,2004MNRAS.350..829P,2006KlahrKley,2007Oishi_etal}. Calculation of radiative transfer, even without iterating for self-consistency, is very computationally intensive. To do this iteratively and coupled with three-dimensional hydrodynamics is challenging, but is the next logical step to improving the accuracy of our results." }, "0801/0801.1491_arXiv.txt": { "abstract": "In this paper, we present results from the complete set of cosmic microwave background (CMB) radiation temperature anisotropy observations made with the Arcminute Cosmology Bolometer Array Receiver (ACBAR) operating at $150\\,$GHz. We include new data from the final 2005 observing season, expanding the number of detector-hours by 210\\% and the sky coverage by 490\\% over that used for the previous ACBAR release. As a result, the band-power uncertainties have been reduced by more than a factor of two on angular scales encompassing the third to fifth acoustic peaks as well as the damping tail of the CMB power spectrum. The calibration uncertainty has been reduced from 6\\% to 2.1\\% in temperature through a direct comparison of the CMB anisotropy measured by ACBAR with that of the dipole-calibrated WMAP5 experiment. The measured power spectrum is consistent with a spatially flat, $\\Lambda$CDM cosmological model. We include the effects of weak lensing in the power spectrum model computations and find that this significantly improves the fits of the models to the combined ACBAR+WMAP5 power spectrum. The preferred strength of the lensing is consistent with theoretical expectations. On fine angular scales, there is weak evidence ($1.1\\sigma$) for excess power above the level expected from primary anisotropies. We expect any excess power to be dominated by the combination of emission from dusty protogalaxies and the Sunyaev-Zel'dovich effect (SZE). However, the excess observed by ACBAR is significantly smaller than the excess power at $\\ell >2000$ reported by the CBI experiment operating at $30\\,$GHz. Therefore, while it is unlikely that the CBI excess has a primordial origin; the combined ACBAR and CBI results are consistent with the source of the CBI excess being either the SZE or radio source contamination. ", "introduction": "\\label{sec:intro} Observations of the cosmic microwave background (CMB) are among the most powerful and important tests of cosmological theory. Measurements of the angular power spectrum of CMB temperature anisotropies on angular scales $> 10^\\prime$ - corresponding to multipoles $\\ell \\lesssim 1000$ - \\citep{spergel06} in conjunction with other cosmological probes \\citep{burles01,cole05,tegmark06,riess07} have produced compelling evidence for the $\\Lambda$CDM cosmological model. At higher multipoles, measurements probe the Silk damping tail of the power spectrum and provide an independent check of the cosmological model. At smaller angular scales, the primary CMB anisotropies originating at redshift z = 1100 are exponentially damped by photon diffusion. This effect, known as Silk damping, makes secondary anisotropies - those induced along the line of sight at lower redshift - increasingly important at higher $\\ell$. At 150$\\,$GHz, for example, the Sunyaev-Zel'dovich effect (SZE) is expected to be brighter than the primary CMB anisotropy at $\\ell \\gtrsim 2500$. The amplitude of the SZE depends sensitively on the amplitude of the matter perturbations, scaling as $\\sigma_8^7$. Measurements of the CMB power spectrum with sufficient sensitivity on arcminute scales not only extend tests of the $\\Lambda$CDM model's ability to accurately predict the features in the power spectrum of primary CMB anisotropy, but also probe the epoch of cluster formation and provide an independent measure of $\\sigma_8$. In this paper, we present the complete results of observations of CMB temperature anisotropies at 150 GHz with 5$^\\prime$ resolution from the Arcminute Cosmology Bolometer Array Receiver (ACBAR) experiment at the South Pole station. Previous measurements of the CMB power spectrum by ACBAR have been presented in \\cite{kuo04} (hereafter K04) and \\cite{kuo07} (hereafter K07). In addition, the angular power spectrum on these angular scales has been measured at 30 GHz by CBI \\citep{readhead04}, VSA \\citep{dickinson04}, and BIMA \\citep{dawson06}, and at 100 and 150 GHz by QuAD \\citep{quad07}. To date, measurements at angular scales $<10^\\prime$ have been consistent with predictions of the primary anisotropy based on measurements at larger angular scales, with one exception. Both CBI~\\citep{mason03,bond05} and BIMA~\\citep{dawson06} observe excess power for $\\ell > 2000$ at 30 GHz compared to the predictions of the $\\Lambda$CDM model. This excess can be explained by the SZE if $\\sigma_8 \\approx 1$, but this value is in tension with the best-fit WMAP5 value of $\\sigma_8 \\approx 0.8$. In K07, we found that while the frequency dependence of the excess is consistent with the SZE, the ACBAR and CBI data could not be used to rule out radio source contamination or systematic errors as the source of the CBI excess. Careful measurements over a broad range of frequencies and angular scales are needed to provide a definitive answer. Current estimates of the primordial power spectrum are consistent with the predictions of slow-roll inflation for a nearly scale-invariant spectrum which may also include a small running of the spectral index. Sparked by the modest evidence for negative running in the WMAP first-year data, a number of authors have investigated how existing data sets limit the allowed inflationary scenarios \\citep{peiris03,mukherjee03,bridle03,leach03}. Small-scale data extend the range over which the primordial power spectrum is measured and can potentially yield information about the mechanism of inflation. This is the third and final ACBAR power spectrum release. The first release in K04 analyzed two fields from the 2001 and 2002 seasons with a conservative field differencing algorithm. The second ACBAR power spectrum, presented by K07, added two more fields from the 2002 season and implemented an improved, undifferenced Lead-Main-Trail (no-LMT) analysis of the dataset. The results presented here improve on the previous work in two ways. First, we include seven additional fields observed in the 2005 Austral winter. These fields double the total number of detector hours and substantially improve the precision of the band-power estimates. In particular, the new fields were selected to dramatically expand ACBAR's sky coverage in order to reduce the cosmic variance contribution to the uncertainty and to improve the multipole resolution on angular scales below $\\ell \\lesssim 1800$. This angular range covers the third to fifth acoustic peaks, making it especially interesting for constraining cosmological models. Second, we implement a new temperature calibration based on a comparison of CMB fluctuations as measured by ACBAR and the WMAP satellite~\\citep{hinshaw08}. This improved calibration tightens constraints on cosmological models found from the combination of high-$\\ell$ ACBAR band-powers with low-$\\ell$ results from other experiments. This paper is organized as follows. In \\S~\\ref{sec:instrument} we review the ACBAR instrument and the CMB observation program. The analysis algorithm is explained in \\S~\\ref{sec:analysis}. Section \\S~\\ref{sec:calib} is an overview of the calibration; the details of cross-calibration between WMAP5 and ACBAR are discussed in Appendix A. Information on ACBAR's beams can be found in \\S~\\ref{sec:beam}. Systematic tests and foreground contamination are discussed in \\S~\\ref{sec:sys}. We present the band-power results in \\S~\\ref{sec:results}, including a discussion of the scientific interpretation. The ACBAR band-powers are combined with the results of other experiments to place constraints on the parameters of cosmological models in \\S~\\ref{sec:parameters}. In \\S~\\ref{sec:conclusion}, we summarize the main results of this paper. ", "conclusions": "" }, "0801/0801.3341_arXiv.txt": { "abstract": "Our \\xmm\\ spectrum of the giant, high-redshift ($z=1.88$) radio galaxy \\qc\\ shows that it contains one of the most powerful, high-redshift, Compton-thick quasars known. Its spectrum is very hard above 2\\kev. The steep \\xmm\\ spectrum below that energy is shown to be due to extended emission from the radio bridge using \\chandra\\ data. The nucleus of \\qc\\ has a column density of $3.5^{+1.4}_{-0.4}\\times 10^{24}$\\pcmsq\\ and absorption-corrected X-ray luminosity of $1.7^{+0.9}_{-0.1}\\times 10^{45}$\\ergps\\ in the $2-10$\\kev\\ band. A lower redshift active galaxy in the same field, SDSS J090808.36$+$394313.6, may also be Compton-thick. ", "introduction": "The powerful, classical double \\frii\\ \\citep{FR} radio galaxy \\qc\\ is unusual as it is one of the highest redshift ($z=1.88$) giant radio galaxies known, spanning $111$\\as\\ on the sky \\citep{lawgreen95} corresponding to a projected size of $945$\\kpc\\ in the cosmology assumed in this paper. Radio galaxies have long been associated with obscured, (optically defined) Type 2 quasars, because they are considered to be the same objects as radio-loud quasars seen through an obscuring torus (e.g. \\citealt{barthel89} and \\citealt{antonucci93}). X-ray observations are particularly efficient at detecting obscured AGN as they provide the observer with several diagnostics, such as the photoelectric cut-off, the iron line equivalent width (EW) and the X-ray to optical or [O III] $\\lambda 5007$ ratio (e.g. \\citealt{bassani99} or, for a review, \\citealt{comastri04}). Sources with an obscuring column greater than $1.5 \\times 10^{24}$\\pcmsq, at which the Thomson scattering optical depth reaches unity, are known as Compton-thick AGN. It is non-trivial to calculate the column density of such sources as there is little (see \\citealt{wilmanfabian99}) direct emission below 10\\kev\\ meaning that analysis relies on interpreting the reflected spectrum at lower X-ray energies. \\suzaku\\ is starting to provide a window into the X-ray regime above 10\\kev\\ enabling direct detection of nuclear emission in the brightest Compton-thick sources such as NGC\\,5728 \\citep{comastri07}. Such sources are important since the X-ray background gives strong evidence for a significant population of Compton-thick AGN (e.g. \\citealt{comastri04}, \\citealt{worsley05}). In this paper, we present recent \\xmm\\ observations of \\qc\\ (which is located at RA 09h08m16.9s, Dec +39d43m26s in J2000). Throughout this paper, all errors are quoted at $1\\sigma$ unless otherwise stated and the assumed cosmology is $\\rm H_{\\rm 0} = 71$\\kmpspmpc, $\\Omega_{0}=1$ and $\\Lambda_{0} = 0.73$. One arcsecond represents $8.518$\\kpc\\ on the plane of the sky at the redshift ($z=1.883\\pm 0.003$) of \\qc\\ and the Galactic absorption along the line-of-sight is taken to be $1.91 \\times 10^{20}$\\pcmsq\\ \\citep{dickeylockman90}. ", "conclusions": "\\label{sec:conclude} Our \\xmm\\ spectrum of \\qc, which is one of the highest redshift giant radio galaxies known, shows it to contain a Compton-thick quasar. \\qc\\ is one of very few such sources currently known and one of even fewer at high redshift. It is likely to be a high-Eddington source with a bolometric luminosity of $\\sim 10^{47}$\\ergps, making it one of the most powerful sources known." }, "0801/0801.0204_arXiv.txt": { "abstract": "% We are conducting a deep near-infrared (NIR) imaging survey of young embedded clusters in the extreme outer Galaxy (hereafter EOG), at the Galactic radius ($R_{\\rm g}$) of more than 18 kpc. The EOG is an excellent laboratory to study the nature of the IMF in a low-metallicity environment with a great advantage of the proximity compared to nearby dwarf galaxies, such as LMC \\& SMC. As a first step, we obtained deep NIR images of Digel Cloud 2 clusters at $R_{\\rm g} \\simeq 19$ kpc using the Subaru 8.2-m telescope. The observed $K$-band luminosity function shows that {\\it IMF in the low metallicity environment down to $\\sim 0.1$ $M_\\odot$ is not significantly different from the typical IMFs in the field and in the nearby star clusters} as was suggested in our earlier work \\citep{Yasui2006}. ", "introduction": "The initial mass function (IMF) in nearby galaxies has been extensively studied with high-resolution imaging using {\\it HST} \\citep[e.g.,][]{{Sirianni2000},{Gouliermis2005}}. It is critical to reach to the substellar mass regime for the complete census of the IMF, because {\\it a characteristic peak mass, which is important to determine the shape of the IMF, is at around $\\sim$ 0.5 $M_\\odot$} \\citep{AnnualReview}. However it is difficult to reach this limit even for the nearest galaxies with active star formation, such as LMC \\& SMC, because of the lack of spatial resolution. Also, observations of embedded clusters at NIR wavelengths are more ideal for investigating the ``true'' IMF right after the cluster formation without being hampered by dust extinction \\citep{AnnualReview}. Therefore, we started a program to study the IMF of the young embedded clusters in the EOG defined here as the region at $R_{\\rm g}$ greater than 18 kpc. The heliocentric distance to EOG is as close as 10 kpc, which is five times closer than LMC ($D \\sim 50$ kpc). The EOG has a similar environment as nearby irregular dwarf galaxies (Im) and damped Ly-$\\alpha$ systems (DLAs): low metallicity \\citep[$\\sim - 1$ dex; e.g.,][]{Rudolph2006}, low gas density ($< 0.01$ cm$^{-3}$), and no or little perturbation from the spiral arms. ", "conclusions": "" }, "0801/0801.2037_arXiv.txt": { "abstract": "{We discuss the equatorial imaging benefits that arise from the addition of the 25-metre dish at Chilbolton to the e-MERLIN array. Its inclusion considerably enhances the capabilities of e-MERLIN on and below the equator. This will become particularly important in the era of ALMA and other upcoming southern hemisphere facilities. We present simulated observations of point sources in the equatorial region of the sky which is the target area for many existing sky surveys. We find that the additional baselines created by the inclusion of the Chilbolton dish favourably adjust the beam shape of e-MERLIN to a more compact and circular shape, with significantly reduced sidelobe structure. Putting aside the benefits of increased collecting area, the modified beam shape has implications for more rapidly reaching a given completeness limit for equatorial surveys.} ", "introduction": "The facilities at Chilbolton Observatory in Hampshire, UK, include a fully steerable 25-metre parabolic antenna which is mainly used for meteorological Doppler-polarisation radar. The use of this antenna in the MERLIN array (Thomasson, 1986) has been discussed for many years, and with the provision of a fast fibre link it would become a prime candidate for inclusion in the e-MERLIN\\footnote{\\tt{http://www.merlin.ac.uk/e-merlin/}} array. Figure~\\ref{fig:merlin} is a version of the diagram presented on the online MERLIN user guide\\footnote{\\tt{http://www.merlin.ac.uk/user\\_guide/OnlineMUG/}} which has been modified in order to show the location of the Chilbolton antenna in addition to the locations of the existing MERLIN stations. Antenna numbers for the existing MERLIN stations have also been added. As can be seen, inclusion of the Chilbolton antenna will boost the number of long and intermediate-length baselines, as well as extending the north-south span of the array, facilitating the sampling of a much greater range of spatial frequencies. This is particularly crucial for `snapshot' mode observations, and large-scale survey programmes where on-source time may necessarily be rather brief. In this article we demonstrate the improved $uv$-plane sampling of an e-MERLIN+Chilbolton array for both snapshot and full synthesis equatorial observations. We present simulations demonstrating the response of both arrays to an unpolarized point source, and discuss the implications of the inclusion of the Chilbolton antenna when carrying out equatorial surveys. All observations are simulated using AIPS with standard imaging procedures. All simulations are at L-band, using 100 frequency channels of 5 MHz each to simulate a 1.3 - 1.8 GHz contiguous band, assuming a spectrally flat source. Each antenna has an efficiency of 0.8 and system temperatures in the range 20 - 33~K. The 76-metre Lovell telescope is not included in the simulated array. \\begin{figure} \\centering \\vspace{215pt} \\special{psfile=merlin2.eps hscale=35 vscale=35 hoffset=0 voffset=0} \\caption{Map showing the current MERLIN array, with the location of the Chilbolton Observatory highlighted in grey. \\label{fig:merlin} } \\end{figure} ", "conclusions": "The geographical location of the Chilbolton antenna introduces baselines which complement the existing e-MERLIN array. This is particularly evident in the case of equatorial observations, where the dominant east-west layout of the existing array biases the $uv$ coverage, resulting in strong linear structure in the sidelobes of a synthesised beam which is highly eccentric. Putting aside the increased sensitivity due to the $\\sim$10\\% increase in collecting area with the additional antenna, the synthesised beam is much more compact and circular for equatorial observations, facilitating more efficient detection of characteristic $\\mu$Jy radio sources, reducing the mapping speed by up to a factor of two (see Table~\\ref{tab:maptimes} for favourably oriented sources. The addition of the Chilbolton antenna would significantly enhance the power of e-MERLIN. There are many experiments that would benefit significantly from this enhancement, such as high-resolution radio surveys complementing optical and infrared deep-fields such as those to be undertaken by VISTA, and high-resolution radio observations of galactic and extra-galactic radio sources targetted with ALMA. Both e-MERLIN and ALMA naturally complement each other due to their similar sub-arcsecond resolutions, despite their operating frequencies differing by a factor of $\\sim$100." }, "0801/0801.3630_arXiv.txt": { "abstract": " ", "introduction": "\\label{intro} Ordinary fluids (e.g., gases and liquids) may be bounded by rigid walls which allow particle number conservation, avoiding evaporation. Macroscopical parameters (pressure, density, and temperature) remain uniform within the box, due to its reduced dimensions. On the other hand, astrophysical fluids (e.g., stars and galaxies) may be conceived as bounded by ``gravitational'' walls which violate particle number conservation by allowing evaporation. The macroscopical parameters exhibit gradients because they cannot remain uniform within the ``gravitational'' box, due to its large-scale dimensions. In particular, sufficiently extended celestial objects show at least two distinct components: core-envelope for stars, core-halo for elliptical galaxies, bulge-disk for spiral and lenticular galaxies, baryonic-nonbaryonic for virialized (matter) density perturbations, and matter-dark energy for virialized (matter + dark energy) density perturbations. On this basis, an investigation on two-component astrophysical fluids appears useful for the comprehension and the interpretation of what is inferred from observations. To this aim, the choice of the density profiles is of basic importance. The laws of ideal and real gases were deduced from ordinary fluids, characterized by uniform density profiles. Accordingly, it is expected that astrophysical fluid laws are related to the specified density profiles, and different laws hold for different matter distributions. Strictly speaking, a density profile should be deduced from the distribution function, or vice versa. Unfortunately, the determination of the distribution function is much more difficult than in one-component systems, and only a few cases have been studied in detail at present (e.g., Ciotti, 1996, 1999). On the other hand, global properties exhibited by simple density profiles (with somewhere negative distribution function) are expected to maintain a similar trend in dealing with much more complex density profiles (with nonnegative distribution function). As the current attempt is mainly aimed to explore global properties instead of local properties, density profiles shall be selected according to their intrinsic simplicity, regardless from the physical meaning of the distribution function. Configurations described by simple density profiles could be sufficiently close to their counterparts described by self-consistent density profiles, and related results hold to a first extent. In any case, the self-consistency of density profiles with respect to nonnegativity of the distribution function, can be checked using a specific theorem (Ciotti and Pellegrini, 1992). To a first extent, the particle number shall be supposed to be conserved, which is equivalent to conceive the boundary of each subsystem as a perfectly reflecting surface in order to avoid evaporation. In this view, a possible choice of macroscopical parameters is: the fractional virial potential energy, $\\phi$; the fractional truncation radius, $y$; and the fractional mass, $m$; as done in a pioneering paper with respect to uniform density profiles i.e. homogeneous configurations (Caimmi and Secco, 1990, hereafter quoted as CS90). Accordingly, each subsystem is supposed to be virialized, in the sense that the virial equations are satisfied by averaging over a sufficiently long time, and particles move within a bounded region (e.g., Landau and Lifchitz, 1966, Chap.\\,II, \\S10; Caimmi, 2007a). Then virial equilibrium is a necessary (but not sufficient) condition for dynamical or hydrostatic equilibrium, which, on the other hand, does not imply pressureless configurations, as the stress tensor is related to the kinetic-energy tensor (e.g., Binney and Tremaine, 1987, Chap.\\,4, \\S2). For sake of simplicity, the applications of the general theory shall be restricted to homeoidally striated density profiles (e.g., Roberts, 1962; Caimmi, 1993; Caimmi and Marmo, 2003, hereafter quoted as CM03). The larger effects of asphericity are expected to occurr in homogeneous configurations, which have widely been investigated (Brosche et al., 1983; Caimmi et al., 1984; CS90; Caimmi and Secco, 1992). Focaloidally striated density profiles involve far larger difficulty (e.g., Caimmi, 1995, 2003). The current investigation is mainly devoted to the following points: (i) expression of an equation of state for two-component systems; (ii) description of global properties deduced from selected density profiles; and (iii) application to elliptical galaxies belonging to a restricted sample, to be represented on the $({\\sf O}y\\phi)$ plane, for fiducial values of model parameters. The work is organized as follows. The basic theory of two-component systems with homeoidally striated density profiles, is reviewed and extended (to include both cored and cuspy matter distributions) in Section \\ref{bath}. The particularization to selected density profiles, involving explicit expressions, is made in Section \\ref{spec}. The results and related global properties are described and discussed in Section \\ref{resu}. Elliptical galaxies belonging to a restricted sample, are represented on the $({\\sf O}y\\phi)$ plane, for fiducial values of model parameters, in Section \\ref{appl}, where some considerations are drawn. The concluding remarks are reported in Section \\ref{conc}. ", "conclusions": "\\label{conc} Two-component systems have been conceived as (two-component) macrogases, and the related equation of state has been formulated using the virial theorem for subsystems (Limber, 1959; Brosche et al., 1983; Caimmi et al., 1984; Caimmi and Secco, 1992), under the restriction of (i) homeoidally striated ellipsoids (Roberts, 1962) and (ii) similar and similarly placed boundaries. Explicit calculations have been performed for a useful reference case and a few cases of astrophysical interest, both in presence and in absence of truncation radius. More specifically, the following cases have been dealt with: IJ= UU, PP, HH, HP, HN, where I and J denote the inner and the outer density profile, respectively, and the other captions relate to the following density profiles: U ($\\rho=$const), P (Plummer, 1911), H (Hernquist, 1990), N (Navarro et al., 1995, 1996, 1997). Shallower density profiles (UU, PP), have been found to yield an equation of state, $\\phi=\\phi(y,m)$, characterized by the occurrence of two extremum points, one maximum and one minimum, as in an earlier attempt (CS90). Steeper density profiles (HH, HP, HN), have been found to produce a similar equation of state where, in addition, a single horizontal inflexion point occurs in a critical isofractional mass curve, and isofractional mass curves related to lower values, $m=M_j/M_i 10$) on time scales of years. These are X-ray binaries and SSS.\\@ Among the active galactic nuclei (AGNs) narrow-line Seyfert 1 galaxies and BL Lac objects show the strongest variabilities. However only a few narrow-line Seyfert 1 galaxies are known in the entire sky, which show flux variability factors larger than 10 on time scales of half a year up to several years. Hence it is very unlikely that one of the strongly variable sources in \\m31\\ would be an AGN. In this paper we report a search for new X-ray sources in the \\xmm\\ observations to the \\m31\\ centre to extend the source catalogue of PFH2005 and a time variability analysis of all the \\m31\\ centre sources. In Sect.~2 information about the used observations and accomplished analysis is provided. Sect.~3 describes the source catalogue extension. The results of the temporal variability analysis are discussed in Sect.~4. Discussion of the individual source classes, including X-ray identifications, are provided in Sect.~5. We draw our conclusions in Sect.~6. ", "conclusions": "In this paper we present an updated source list of the central area of the bright Local Group spiral galaxy \\m31, using the observations from June 2000 to July 2004 available from the \\xmm\\ archive. We extended the source catalogue by PFH2005, based on the merged images of the observations from 2000 to 2002 by searching sources in the observations of 2004 and reexamining the observations used in PFH2005 individually. To classify or identify more of the sources, we examined their long term time variability. We obtained 39 sources in addition to the 265 reported by PFH2005 in the field. The identification and classification of these sources is based on properties in the X-ray wavelength regime: hardness ratios and temporal variability. In addition, information from cross correlations with \\m31\\ catalogues in the radio, infra-red, optical and X-ray wavelength regimes are used. We detected three SSS candidates, one SNR and six SNR candidates, one GlC candidate, three XRBs and four XRB candidates. Additionally we identified one foreground star candidate and classified fifteen sources as $<$hard$>$, which may either be XRBs or Crab-like SNRs in \\m31\\ or background AGNs. The remaining five sources remain unidentified and without classification. Two sources were found to be extended. One of them was classified as $<$hard$>$. The other stays without classification. To examine the time variability we calculated the flux or at least an upper limit at the source positions in each observation. We determined the variability factor and significance parameter for each source, comparing the XID flux ratios of the different observations with each other. The time variability helped us to decide if a source classified as $<$hard$>$ in PFH2005 can be an XRB candidate. In addition we could use time variability to distinguish between foreground star and SNR candidates. Six sources of PFH2005, which were classified as $<$hard$>$, show distinct time variability. Based on that variability, their hardness ratios and the strong absorption in the centre of \\m31\\ we suggest these sources as XRB candidates. The SNR classification from source 295 was changed to foreground star due to the distinct time variability we found and its identification with a faint stellar object. Other SNR classifications (sources 316, 318) were rejected due to time variability of the sources. To verify our suggested classifications further investigations, including at other wavelengths will be necessary." }, "0801/0801.1403_arXiv.txt": { "abstract": "We explore the importance of magnetic-field-oriented thermal conduction in the interaction of supernova remnant (SNR) shocks with radiative gas clouds and in determining the mass and energy exchange between the clouds and the hot surrounding medium. We perform 2.5D MHD simulations of a shock impacting on an isolated gas cloud, including anisotropic thermal conduction and radiative cooling; we consider the representative case of a Mach 50 shock impacting on a cloud ten-fold denser than the ambient medium. We consider different configurations of the ambient magnetic field and compare MHD models with or without the thermal conduction. The efficiency of the thermal conduction in the presence of magnetic field is, in general, reduced with respect to the unmagnetized case. The reduction factor strongly depends on the initial magnetic field orientation, and it is minimum when the magnetic field is initially aligned with the direction of shock propagation. The thermal conduction contributes to suppress hydrodynamic instabilities, reducing the mass mixing of the cloud and preserving the cloud from complete fragmentation. Depending on the magnetic field orientation, the heat conduction may determine a significant energy exchange between the cloud and the hot surrounding medium which, while remaining always at levels less than those in the unmagnetized case, leads to a progressive heating and evaporation of the cloud. This additional heating may contrast the radiative cooling of some parts of the cloud, preventing the onset of thermal instabilities. ", "introduction": "\\label{sec1} The interaction of the shock waves of supernova remnants (SNRs) with the magnetized and inhomogeneous interstellar medium (ISM) is responsible of the great morphological complexity of SNRs and certainly plays a major role in determining the exchange of mass, momentum, and energy between diffuse hot plasma and dense clouds or clumps. These exchanges may occur through, for example, hydrodynamic ablation and thermal conduction and, among other things, lead to the cloud crushing and to the reduction of the Jeans mass causing star formation. The propagation of hot SNR shock fronts in the ISM and their interaction with local over-dense gas clouds have been investigated with detailed hydrodynamic and MHD modeling. The most complete review of this problem in the unmagnetized, non-conducting, and non-radiative limits is provided by \\cite{1994ApJ...420..213K}. These studies have shown that the cloud is disrupted by the action of both Kelvin-Helmholtz (KH) and Rayleigh-Taylor (RT) instabilities after several crushing times, with the cloud material expanding and diffusing into the ambient medium. An ambient magnetic field can both act as a confinement mechanism of the plasma and be modified by the interstellar flow and by local field stretching. Also, a strong magnetic field is known to limit hydrodynamic instabilities developing during the shock-cloud interaction by providing an additional tension at the interface between the cloud and the surrounding medium \\citep[e.g][]{1994ApJ...433..757M, 1996ApJ...473..365J}. The interaction of the shock with a {\\it radiative} cloud has been only recently analyzed in detail \\citep[e.g.][]{2002A&A...395L..13M, 2004ApJ...604...74F}. 2D calculations have shown that the effect of the radiative cooling is to break up the clouds into numerous dense and cold fragments that survive for many dynamical timescales. In the case of the interaction between magnetized shocks and radiative clouds, the magnetic field may enhance the efficiency of the radiative cooling, influencing the size and distribution of condensed cooled fragments \\citep{2005ApJ...619..327F}. The role played by the thermal conduction during the shock-cloud interaction has been less studied so far. In a previous paper, \\cite{2005A&A...444..505O} (hereafter Paper I) have addressed this point in the unmagnetized limit. In particular, we have investigated the effect of thermal conduction and radiative cooling on the cloud evolution and on the mass and energy exchange between the cloud and the surrounding medium; we have selected and explored two different physical regimes chosen so that either one of the processes is dominant. In the case dominated by the radiative losses, we have found that the shocked cloud fragments into cold, dense, and compact filaments surrounded by a hot corona which is ablated by the thermal conduction. Instead, in the case dominated by thermal conduction, the shocked cloud evaporates in a few dynamical timescales. In both cases, we have found that the thermal conduction is very effective in suppressing the hydrodynamic instabilities that would develop at the cloud boundaries, preserving the cloud from complete destruction. \\cite{orlando2} and \\cite{2006A&A...458..213M} have studied the observable effects of thermal conduction on the evolution of the shocked cloud in the X-ray band. Here, we extend the previous studies by investigating the effect of the thermal conduction in a magnetized medium, unexplored so far. Of special interest to us is to investigate the role of anisotropic thermal conduction - funneled by locally organized magnetic fields - in the mass and energy exchange between ISM phases. In particular, we aim at addressing the following questions: How and under which physical conditions does the magnetic-field-oriented thermal conduction influence the evolution of the shocked cloud? How do the mass mixing of the cloud material and the energy exchange between the cloud and the surrounding medium depend on the orientation and strength of the magnetic field and on the efficiency of the thermal conduction? To answer these questions, we take as representative the model case of a shock with Mach number $\\mach = 50$ (corresponding to a post-shock temperature $T\\approx 4.7\\times 10^6$ K for an unperturbed medium with $T=10^4$ K) impacting on an isolated cloud ten-fold denser than the ambient medium. Paper I has shown that, in this case, the thermal conduction dominates the evolution of the shocked cloud in the absence of magnetic field. Around this basic configuration, we perform a set of MHD simulations, with different interstellar magnetic field orientations, and compare models calculated with thermal conduction turned either ``on'' or ``off'' in order to identify its effects on the cloud evolution. The paper is organized as follows: in Sect. \\ref{sec2} we describe the MHD model and the numerical setup; in Sect. \\ref{sec3} we discuss the results; and finally in Sect. \\ref{sec4} we draw our conclusions. ", "conclusions": " \\begin{enumerate} \\item In general, we found that the effects of thermal conduction on the evolution of the shocked cloud are reduced in the presence of an ambient magnetic field with respect to the unmagnetized cases investigated in Paper I. The efficiency of anisotropic thermal conduction strongly depends on the initial magnetic field orientation and configuration. This efficiency is the largest when the initial {\\mag} is aligned with the direction of propagation of the shock front, and is the smallest when {\\mag} is aligned with the cylindrical cloud, namely when the heat conduction is completely suppressed by the magnetic field. \\item We found that the hydrodynamic instabilities are suppressed efficiently by the anisotropic thermal conduction when the initial magnetic field is perpendicular to the cylindrical cloud (a configuration referred to as ``external fields''). On the contrary, in the case of {\\mag} parallel to the cylindrical axis of the cloud (i.e. when the field has component only along the $z$ axis - internal field), hydrodynamic instabilities develop at the cloud boundary. We found that, for the parameters of the simulations chosen, the magnetic tension is unable to suppress alone the hydrodynamic instabilities. \\item As for thermal instabilities, we found that, depending on the magnetic field orientation, the heat flux contributes to the heating of some parts of the cloud, reducing the efficiency of radiative cooling there, and preventing any thermal instability. \\item The mass loss of the cloud due to mixing with the surrounding medium is mainly driven by hydrodynamic instabilities; in the case of external fields (initial {\\mag} perpendicular to the cylindrical cloud) the anisotropic thermal conduction reduces the mass mixing of the cloud. In any case, the mass loss rate is larger than that in the corresponding unmagnetized case ($\\dot{m}_{\\rm cl}\\approx 1.5\\times 10^{-7} L_{\\rm pc}~~M_{\\odot}$~yr$^{-1}$, i.e. $\\sim 5$\\% of the cloud mass is in mixed zones at $t = 3.5\\;\\tau_{\\rm cc}$), but can get very high when the thermal conduction is completely suppressed ($\\dot{m}_{\\rm cl}\\approx 4\\times 10^{-6} L_{\\rm pc}~~M_{\\odot}$~yr$^{-1}$, i.e. $\\sim 45$\\% of the cloud mass is in mixed zones at $t = 3.5\\;\\tau_{\\rm cc}$). \\item The thermal conduction mostly rules the energy exchange between the cloud and surrounding medium. The exchange is favored when the magnetic field configuration is such that the conductive flow is not suppressed (i.e. external field configurations, $B_{\\rm x}$ and $B_{\\rm y}$ cases), but it is never as high as in the absence of magnetic field. In the $B_{\\rm y}$ case, the cloud core is efficiently heated and evaporates in few dynamical timescales. \\item In general, the initial magnetic field strength has a small influence on the dynamic and thermal evolution of the shocked cloud for the ranges of values explored in this paper (namely $0.26~\\mu\\mbox{G} \\leq |\\mag| \\leq 2.63~\\mu\\mbox{G}$). \\end{enumerate} It is worth noting that some details of our simulations depend on the choice of the model parameters. For instance, the onset of thermal instabilities or the evaporation of the whole cloud depends on the initial shock Mach number, and on the density and dimensions of the cloud. The cases that we present here (i.e. $\\mach = 50$, $\\chi=10$, and different configurations of \\mag) are representative of a regime in which both the thermal conduction and the radiative cooling play an important role in the evolution of the shocked cloud. Nevertheless, our analysis proves that anisotropic thermal conduction can not be neglected in investigations of the evolution of shocked interstellar clouds. In our simulations, we consider laminar thermal conduction, although regions of strong turbulence of different strength and extent develop in the system (for instance, at the shear layers along the cloud boundary or at the vortex sheets in the cloud wake). In fact, the turbulence in these regions may have a significant effect on thermal conduction, leading to significant deviations of thermal conductivity from its laminar values (e.g. \\citealt{2001ApJ...562L.129N}; \\citealt{2006ApJ...645L..25L}); in some cases, the turbulence may enhance the heat transfer, exceeding the classical Spitzer value (\\citealt{2006ApJ...645L..25L}). As a result, thermal conduction may be not only anisotropic (in the presence of the magnetic field) but also ``inhomogeneous'' due to the presence of turbulence. However, even modeling accurately the turbulent thermal conductivity, we do not expect significant changes in the results of our $B_{\\rm z}$ case, being the thermal conduction strongly ineffective in the whole spatial domain; in the remaining cases ($B_{\\rm x}$ and $B_{\\rm y}$), our modeled thermal conductivity could be underestimated in regions of strong turbulence, affecting some details of the simulations but not the main conclusion of the paper that, in general, anisotropic thermal conduction can play an important role in the evolution of the shocked cloud. Note also that the field configurations studied in this work are highly idealized. More realistic fields are expected to have more complex topologies and, often, the field can be tangled and chaotic. In the latter case, the thermal conduction will approach isotropy, whereas the effect of MHD turbulence is expected to partially suppress the heat transfer within a factor $\\sim 5$ below the classical Spitzer estimate\\footnote{As already discussed, the MHD turbulence can even enhance the heat transfer in some cases (see \\citealt{2006ApJ...645L..25L}).} (\\citealt{2001ApJ...562L.129N}; \\citealt{2006ApJ...645L..25L}). The shock-cloud collision in the presence of an organized ambient magnetic field, discussed here, and that in the absence of magnetic field can be considered as extreme cases: the former leading to highly anisotropic thermal conduction, the latter to the classical Spitzer thermal conduction. The case of chaotic magnetic field is expected to fall in between these two. Our simulations were carried out in 2.5D Cartesian geometry, implying that the modeled clouds are elongated along the $z$ axis. This choice is expected to affect some details of the simulations but not our main conclusions. Adopting a 3D Cartesian geometry and modeling a spherical cloud, the highly symmetric shock transmitted into the cloud converging on the symmetry axis would lead to compression stronger than those found in our 2.5D simulations, enhancing the radiative cooling. Also, 3D simulations would provide an additional degree of freedom for hydrodynamic instabilities, increasing the mass loss rate of the cloud in the cases in which the mass mixing of cloud material is driven by instabilities. Note that, for a spherical cloud, our $B_{\\rm x}$ and $B_{\\rm z}$ cases no longer differ. Finally we assume, in our simulations, that the cloud and the ambient material have the same composition, implying that microscopic mass mixing due to shear instabilities would be irrelevant. In a more realistic condition, a cold dense cloud may have a different composition from the hot ambient flow and the degree of microscopic mixing may translate into different spectral signatures of the system. In this case, species diffusion could also be important, along with thermal conduction, to determine the degree of microscopic mixing of the materials and, consequently, one would have to ask about the typical values of the Lewis number (i.e. the ratio of thermal diffusivity to mass diffusivity) in the system. It is worth emphasizing that the quantitative results of our simulations depend on the physical parameters of the model (shock Mach number, density contrast and dimension of the cloud, etc.) as well as on the basic assumptions of the model (geometry of the cloud, geometry of the ambient magnetic field, laminar thermal conduction, composition of the cloud and of the ambient medium, etc.). Nevertheless, our results undoubtedly show that the magnetic-field-oriented thermal conduction can play an important role in the evolution of the shock-cloud interaction (which depends on the magnetic field orientation and configuration) and, in particular, in the mass and energy exchange between the cloud and the hot surrounding medium. We conclude, therefore, that a self-consistent and quantitative description of the interaction between magnetized shock-waves and interstellar gas clouds should include the effects of thermal conduction. The results presented here are interesting for the study of middle-aged SNR shells expanding into a magnetized ISM and whose morphology is affected by ISM inhomogeneities (for instance, G272.2-3.2, e.g. \\citealt{1996rftu.proc..247E}; Cygnus Loop, e.g. \\citealt{2002AJ....124.2118P}; Vela SNR, e.g. \\citealt{2005A&A...442..513M}). It will be further interesting to extend the present study, by modeling the shock-cloud interaction in 3D with radiative cooling, anisotropic thermal conduction, and magnetic field included and, even, considering detailed comparisons of model results with observations." }, "0801/0801.4010_arXiv.txt": { "abstract": "I summarize the highlights of the conference. First I provide a brief history of the beach symposia series our massive star community has been organizing. Then I use most of my allocated space discussing what I believe are the main answered and open questions in the field. Finally I conclude with a perspective of the future of massive star research. ", "introduction": "This is the ninth meeting in what has become the traditional beach symposia series of the massive star community. The series had its origins in 1968 at a workshop on Wolf-Rayet (W-R) stars in Boulder (definitely not a beach location). At that time, the cosmic significance of W-R stars was largely unknown but understanding their nature was deemed important enough for follow-up meetings. This then led to IAU Symp. 49 in Argentina in 1971, and to the subsequent tradition of holding massive star symposia at roughly five-year intervals. For those who like statistics, here is the complete tally: IAU Symp. 83 (1978, Canada), IAU Symp. 99 (1981, Mexico), IAU Symp. 116 (1985, Greece), IAU Symp. 143 (1990, Indonesia), IAU Symp. 163 (1994, Italy), IAU Symp. 193 (1998, Mexico), IAU Symp. 212 (2002, Spain), and finally IAU Symp. 250 (2007, USA). Each meeting had its distinct flavor and theme. For instance, IAU Symp. 163 had the emphasis on binary stars, and IAU Symp. 116 allocated a large fraction of time to massive stars in Local Group galaxies. Overall, a clear trend is apparent: the median distance of the astronomical objects at each meeting has been steadily increasing. Distances were expressed in kpc in the early symposia, then became Mpc, and during this meeting redshifts larger than 3 were commonly mentioned. Those who attended IAU Symp. 49 would have been astounded had they known how the field would have progressed by the time of IAU Symp. 250. Even as recently as during IAU Symp. 193 (which I attended as well), it would have been preposterous to discuss massive stars in connection with gamma-ray bursts, Lyman-break galaxies, and Population III stars. I will finish my historical notes with a bit of trivia. Three participants of this symposium have witnessed this rapid development from the very beginning. They already participated in the 1971 meeting: Peter Conti, Lindsay Smith\\footnote{I apologize to Lindsay Smith whom I failed to name in my original talk.}, and Nolan Walborn. Peter Conti distinguished himself with an impressive feat: he attended {\\em all} nine beach symposia. His record would have been tied were it not for Virpi Niemela's untimely death in 2006. ", "conclusions": "" }, "0801/0801.3295_arXiv.txt": { "abstract": " ", "introduction": "The primordial power spectrum or 2-point correlation function of the temperature anisot\\-ropies in the Cosmic Microwave Background (CMB) is well-measured out to large multipoles. The higher moments of a distribution are, in general, independent of the 2-point function, but the CMB anisotropies are at least approximately Gaussian. Theoretically, we know that non-Gaussian component to the CMB will always be induced by non-linear gravitational couplings between modes after they reenter the horizon \\cite{Pyne:1995bs} while single field slow-roll inflation yields a primordial non-Gaussian component roughly 100 times smaller than that induced by gravitational mode-couplings \\cite{Acquaviva:2002ud,Maldacena:2002vr}. The two terms are additive, so recovering this latter primordial signal is next to impossible, even before contending with cosmic variance. Multi-field models may generate larger non-Gaussianities \\cite{Rigopoulos:2005us,Seery:2005gb,Vernizzi:2006ve,Battefeld:2006sz,Yokoyama:2007uu}, but this is by no means a generic property of these scenarios. Consequently, the detection of a primordial 3-point function would immediately falsify a very large class of inflationary models. Conversely, non-slow-roll models with higher order derivative terms, such as DBI inflation \\cite{Silverstein:2003hf,Alishahiha:2004eh,Chen:2004gc,Chen:2005ad} and k-inflation \\cite{ArmendarizPicon:1999rj,Garriga:1999vw}, do typically generate large non-Gaussianities \\cite{Chen:2006nt}. Further references, include more complicated multi-field, non-local or ghost theories, can be found in Refs.~\\cite{Li:2008qc,Bean:2008ga,Lyth:2002my,Chen:2007gd,Gupta:2002kn,Moss:2007cv,Barnaby:2007yb,ArkaniHamed:2003uz,Cheung:2007st}. In this paper we investigate simple models -- in the sense that they are driven by a single, minimally coupled scalar field with a canonical kinetic term -- which generate substantial non-Gaussianities. Constraining the non-Gaussian signal in a CMB dataset is a highly non-trivial problem. Firstly, it depends on the choice of estimators. At the moment, only two concrete estimators have been constructed: the $f_{NL}^{local}$ and $f_{NL}^{equil}$ forms \\cite{Komatsu:2001rj,Komatsu:2003iq,Babich:2004gb,Spergel:2006hy,Creminelli:2006gc,Creminelli:2006rz,Yadav:2007ny}, and both are scale-invariant. The recent claim of a detection of a non-zero 3-point function \\cite{Yadav:2007yy} in the WMAP 3-year data \\cite{Spergel:2006hy} relies on the estimator developed in Ref.~\\cite{Yadav:2007ny,Creminelli:2006gc,Komatsu:2003iq} which is of the ``local'' form \\cite{Salopek:1990jq,Komatsu:2001rj}. However non-Gaussianities that have strong scale dependence are not well-described by $f_{NL}^{local}$ and $f_{NL}^{equil}$, and will require scale-dependent estimators, based on theoretically motivated ansatzen % for the primordial non-Gaussianities. In addition to computing the primordial 3-point term, comparing the predictions of a specific inflationary model to the CMB requires us to evolve this signal through to recombination, and to project it onto the sphere on the sky, which is a convolution of the primordial 3-point term with 3 $l$-valued spherical Bessel functions. In general, this process is computationally expensive but simplifies dramatically if the 3-point function has special algebraic properties \\cite{Smith:2006ud,Fergusson:2006pr}. Using Maldacena's elegant formalism \\cite{Maldacena:2002vr}, the primordial 3-point curvature correlation function is computed via a set of integrals (over time) of products of three mode functions (or their derivatives) and slow roll parameters. Looking at these integrals, we identify two mechanisms which can create substantial non-Gaussianities. The first class consists of potentials with a localized violation of slow roll. This can take the form of a small localized feature, including the step models \\cite{Adams:2001vc,Komatsu:2003fd} whose 3-point term was first accurately computed by the present authors in \\cite{Chen:2006xj}. In addition models with a short inflationary phase can have initial transients in their dynamics, which will not be fully erased before the longest modes leave the horizon. In these cases, the 3-point term for modes which are leaving the horizon during the violation of slow roll can be magnified by three orders of magnitude, without ruining the fit to the 2-point function. The second class arises when a small ripple is superimposed on top of an otherwise smooth potential. This induces a ``resonance'' inside one of these integrals, giving the 3-point function a substantial amplitude {\\em before} the modes cross the horizon. In the former case the 2-point function may be substantially modified; in the second case the modification of the 2-point function is very small, even though the 3-point function is large. The latter mechanism has not been described previously and the seemingly contrived field theoretic potential may in fact arise naturally in brane inflation models \\cite{Bean:2008na}. With this information in hand, we construct rough analytic approximations to the corresponding 3-point terms. These expressions have the factorizable form required by \\cite{Smith:2006ud}, which means that we can efficiently compute constraints on step potentials or similar models from CMB data, although at this point we are only interested in a qualitative match to the numerically computed 3-point term. % We check our semi-analytic estimates for the 3-point function using an improved version of the code described in \\cite{Chen:2006xj}, which has a much cleaner scheme for removing numerical divergences in the 3-point integrals. The integrands are products of large, rapidly oscillating terms. Analytically, these are finite, but their {\\em numerical\\/} evaluation is non-trivial, and we show how to transform them into an explicitly finite form before the numerical evaluation is carried out. From a practical perspective, this means we can compute the 3-point function for ``triangles'' which contain two very different scales. The paper is organized as follows. In Section \\ref{sect:MW} we review the computation of the 3-point function, and in Section \\ref{sect:NumInt} we describe the numerical algorithm. Sections \\ref{sect:horizon} and \\ref{sect:subhorizon} discuss the two distinct \\emph{mechanisms} for generating large primordial non-Gaussianities within minimally coupled single scalar field inflation, and we derive the approximate analytic ansatzen for the 3-point function. We then use our numerical methods to compute the exact 3-point function for these models, and show that these approximations yield fair representations of the underlying non-Gaussianities. In Section \\ref{sect:Conclusions} we summarize and discuss our results and plans for future work. We follow the notational conventions of \\cite{Chen:2006xj} and set the reduced Planck mass $M_p$ to unity except when presenting final results or defining parameter values. ", "conclusions": "\\label{sect:Conclusions} We studied two distinctly mechanisms for generating large non-Gaussianities within single field inflation. We derive the approximate 3-point correlation functions using semi-analytic methods, which are in the computationally useful \\emph{factorizable} form. In the first mechanism, non-Gaussianity is generated at horizon crossing by either a feature in the potential, or an initial transient in the inflationary dynamics. In the second case, oscillating slow roll parameters induce a resonance which leads to the generation of a non-Gaussian signal well before horizon crossing. In both cases, the 3-point function depends strongly on the individual wavelengths of the modes in the ``triangle''. With a single, sharp feature or non-standard initial conditions, the 3-point is only enhanced in modes which are crossing the horizon as the inflaton traverses the ``feature''. In a resonance model such as (\\ref{Vexample}), the physical non-Gaussianity is present at all scales. In the resonance case, we showed that the 3-point function is periodic with a period $\\Delta K$ proportional to, and smaller than, $K=k_1+ k_2+ k_3$. This non-Gaussianity will peak starting from a fixed scale when projected onto the CMB sky. Assume for simplicity that $K \\sim \\ell$ where $\\ell$ is the CMB multipole, and denote $\\Delta \\ell$ as the oscillation period. At larger scales where the oscillation spanning $\\Delta \\ell <1$, this non-Gaussianity presumably cannot be resolved experimentally. For the numerical example considered here, it would become visible at $\\ell \\sim \\CO(100)$ where $\\Delta \\ell$ starts to exceed $\\CO(1)$, inducing an effective scale-dependence in this signal. With a feature in the potential, the resulting transient violation of slow-roll generically leads to an oscillatory and scale-dependent 3-point function. We wrote down a heuristic and factorizable scale-dependent expression for this signal and showed that it captured the qualitative properties seen in the exact numerical evaluations of the corresponding integrals. The 3-point correlation decays as we move away from $K \\sim k_{*}$, where $k_{*}$ corresponds to the scale leaving the horizon as the inflaton traverses the feature. We also show that while non-standard initial conditions such as the fast-roll model of \\cite{Contaldi:2003zv} can generate large non-Gaussianities relative to our usual expectations for single field inflation, the amplification is not expected to lift them above the ``noise'' of non-linearities produced by post-inflationary gravitational evolution alone. In addition, we have presented detailed description of a general numerical method for computing the 3-point correlation functions of primordial perturbations from canonical single scalar field inflationary models with arbitrary potentials. We show that while the integrals themselves are formally convergent, they need to be regularized as the integrands are oscillatory, and show how this can be accomplished analytically, rendering the numerical integrals rapidly convergent. There is much further work to be done. Our immediate goal \\cite{preparation} is to use our heuristic ansatzen to construct an optimal estimator with which to search for scale-dependent non-Gaussianities in the cosmic microwave background, and estimate the likely bounds that future missions can put on this signal. On the theoretical front, we plan to investigate the details of non-Gaussianities generated by multi-field models \\cite{Seery:2005gb,Battefeld:2006sz} within the horizon, and with non-standard kinetic terms. Moreover, while the analytic approximations we present here capture the qualitative form of the 3-point function, they are not intended to provide a precise quantitative match to the numerically computed values, but these approximations can certainly be improved." }, "0801/0801.2770_arXiv.txt": { "abstract": "Feedback from star formation is thought to play a key role in the formation and evolution of galaxies, but its implementation in cosmological simulations is currently hampered by a lack of numerical resolution. We present and test a sub-grid recipe to model feedback from massive stars in cosmological smoothed particle hydrodynamics simulations. The energy is distributed in kinetic form among the gas particles surrounding recently formed stars. The impact of the feedback is studied using a suite of high-resolution simulations of isolated disc galaxies embedded in dark halos with total mass $10^{10}$ and $10^{12}~\\Msolh$. We focus in particular on the effect of pressure forces on wind particles within the disc, which we turn off temporarily in some of our runs to mimic a recipe that has been widely used in the literature. We find that this popular recipe gives dramatically different results because (ram) pressure forces on expanding superbubbles determine both the structure of the disc and the development of large-scale outflows. Pressure forces exerted by expanding superbubbles puff up the disc, giving the dwarf galaxy an irregular morphology and creating a galactic fountain in the massive galaxy. Hydrodynamic drag within the disc results in a strong increase of the effective mass loading of the wind for the dwarf galaxy, but quenches much of the outflow in the case of the high-mass galaxy. \\endabstract \\keywords methods: numerical --- ISM: bubbles --- ISM: jets and outflows --- galaxies: evolution --- galaxies: formation --- galaxies: ISM \\endkeywords ", "introduction": "Core-collapse supernovae (SNe) and winds from massive stars feed back energy into the interstellar medium (ISM). The energy released by massive stars can efficiently suppress subsequent star formation by destroying dense, star forming clouds, by generating supersonic turbulence and, if the star formation density is sufficiently high for individual supernova remnants to overlap, by blowing gas out of the disc. In starburst galaxies feedback from star formation may result in the development of a galaxy-wide superwind which may (temporarily) remove a large fraction of the gas, while in galaxies with less intense star formation feedback may lead to the development of a galactic fountain. Because these feedback mechanisms operate on time scales that are very short compared to the age of the universe, they can lead to self-regulation. Feedback from star formation is thought to be a crucial ingredient for models of the formation and evolution of galaxies. Without it, star formation becomes much more efficient than observed, particularly in low-mass galaxies \\cite[e.g.][]{White&Frenk1991}. Among other things, galactic winds are also thought to be responsible for the enrichment of the intergalactic medium with heavy elements \\cite[e.g.][]{Aguirre2001} and for pre-heating the gas that ends up in groups and clusters of galaxies \\cite[e.g.][]{Ponman1999}. Modelling SN feedback in simulations of the formation of galaxies is known to be a difficult task, mostly because the injected thermal energy tends to be radiated away well before it has any hydrodynamical effect \\citep[e.g.][]{Katz1996,Balogh2001}. This overcooling problem is probably caused by the fact that current state-of-the-art simulations still lack the resolution to capture the physics of the multiphase ISM \\cite[e.g.][]{Ceverino&Klypin2007}. A typical SN ejects $\\sim 1~\\Msun$ at $\\sim 10^4~\\kms$, which corresponds to a kinetic energy $\\sim 10^{51}~\\erg$. Since the sound-crossing time is initially much smaller than the radiative cooling time, the remnant starts out as an adiabatic blast wave. Once radiative losses become important, the energy-conserving blast wave gives way to a momentum conserving snow-plough phase, whose deceleration is determined mostly by the amount of mass that is swept up by the wind. The initial phases in the evolution of a superbubble driven by multiple SNe are very similar to the evolution of an individual SN remnant. However, if a superbubble blows out of the disc, its subsequent evolution may be strongly affected by the ram pressure of infalling gas and the gravitational attraction of the galaxy \\cite[e.g.][]{Silich1998,Maclow1999,Dercole1999,Silich2001,Fujita2004,Dubois2008}. Current simulations lack the resolution to resolve individual SNe or, in the case of cosmological simulations\\footnote{By cosmological simulations we mean simulations that model the evolution of a representative volume of the universe, as opposed to simulations that zoom in on one or a few galaxies.} even superbubbles. The energy released by dying stars in a single resolution element per time step is typically distributed over a mass that exceeds the mass of SN ejecta by many orders of magnitude. The initial expansion velocity and post-shock temperature are therefore underestimated by large factors. Since the radiative cooling time scales as $t_c \\propto T^{1/2}$ above 1~keV, this implies that the cooling time is greatly underestimated. Because the injection radius is also far too large, the initial cooling time tends to be smaller than the bubble sound-crossing time. Thus, the simulation will skip the adiabatic blast-wave phase and the energy will typically be radiated away before a significant fraction of the thermal energy has been converted into kinetic energy. The fact that poor resolution results in inefficient thermal feedback is generally attributed to a lack of spatial resolution: real SNe explode in hot bubbles of low-density gas, whereas the ISM in cosmological simulations consists of a single dense, ``warm'' phase. Mass resolution is, however, more fundamental. Without sufficient mass resolution, the first SNe will not be able to create a hot, low-density ISM phase in the first place. Cosmological simulations must resort to sub-grid recipes to solve the overcooling problem.\\footnote{In the absence of efficient feedback mechanisms, the amount of cold gas predicted by cosmological simulations is limited by resolution. Hence, it is possible to roughly match the observed amount of cold gas at a particular redshift by tuning the resolution.} Two types of solutions appear to work: injecting (part of) the SN energy in kinetic rather than thermal form \\cite[e.g.][]{Navarro1993,Mihos&Hernquist1994,Kawata2001,Kay2002,Springel2003,Oppenheimer2006,Dubois2008} and/or suppressing radiative cooling by hand \\cite[e.g.][]{Gerritsen1997,Mori1997,Thacker2000,Kay2002,Sommer-Larsen2003,Brook2004,Stinson2006}. Although the suppression of cooling enables the efficient conversion of thermal energy to kinetic energy, the maximum wind velocity will still be underestimated if the total mass of the neighboring resolution elements exceeds that of a superbubble. Kinetic feedback can alleviate this problem by kicking only a small fraction of the resolution elements near the star particle. Kinetic feedback thus gives us the freedom to distribute a fixed amount of kinetic energy over a varying amount of mass. However, using this freedom to increase the initial wind velocity has the drawback that the imposed winds become more poorly sampled and therefore less isotropic. With increasing resolution, the two types of sub-grid models for the generation of galactic winds are expected to converge. A third approach, which is often combined with one of the above, is to employ a sub-grid model to describe the multiphase ISM \\cite[e.g.][]{Yepes1997}. Examples include imposing an effective equation of state which specifies the total pressure of the ISM as a function of its mean density \\cite[e.g.][]{Springel2003} and, at least for the case of smoothed particle hydrodynamics simulations, explictly decoupling thermal phases by using different types of particles to represent the hot and cold gas phases \\cite[e.g.][]{Marri&White2003,Scannapieco2006}. While galactic winds can in principle be triggered naturally if the latter method is used, this again requires ad-hoc solutions in the absence of sufficient resolution. To prevent the overcooling problem, simulations that impose an effective equation of state for the ISM must either make it extremely stiff, resulting in discs that are much thicker and smoother than observed, or employ a sub-grid recipe for galactic winds. However, even if it does not directly generate winds, the use of an effective equation of state for dense gas can be considered a necessary ingredient for simulations that lack the resolution and/or physics to model the multiphase ISM. If the equation of state is not modified by hand, gas will accumulate at unrealistically low temperatures and high densities in the absence of resolved feedback processes. As we discussed in \\cite{Schaye2008}, using a power-law equation of state with a polytropic index of $4/3$ has the advantage of yielding a Jeans mass that is independent of the density which makes it possible to suppress spurious fragmentation due to a lack of resolution. It is, however, important to note that maintaining an effective equation of state in the presence of radiative losses would, in reality, require energy. The amount of energy that is required depends on unresolved physical processes and can therefore not be reliably determined. Imposing an equation of state for dense gas therefore implies that the energy available for any wind sub-grid model must be less than the energy provided by star formation. The most widely used recipe for galactic winds in SPH simulations is the kinetic feedback model implemented by \\cite{Springel2003} (hereafter SH03). Motivated by the desire to impose the net mass loading and velocity of the wind after it has escaped from the disc and by the wish to make the recipe insensitive to numerical resolution, hydrodynamical forces on the wind particles are temporarily switched off in the SH03 model. Thus, winds cannot blow bubbles in the disc, drive turbulence or create channels in gas with densities typical of the ISM. Their sole effect on the disc is the removal of fuel for star formation by the ejection of wind particles. Another aspect of the SH03 recipe is that wind particles are selected stochastically from all the dense (i.e.\\ star-forming) particles in the simulation and are therefore not constrained to be neighbours of newly-formed stars. While this non-local feedback solves some numerical problems, we will argue that it can lead to undesirable behavior in galaxies that do not contain large numbers of particles (i.e.\\ most galaxies in cosmological simulations), particularly if metal enrichment is included. We present and test a modified a variation of the SH03 recipe in which the winds are local and not decoupled hydrodynamically. In this paper we will focus on the effects of hydrodynamical drag within the ISM. Using high-resolution SPH simulations of isolated disc galaxies we show that pressure forces exerted by and on wind particles have a dramatic effect on both the structure of the ISM and the development of galactic winds. Hydrodynamical drag on wind particles converts low-mass disc galaxies into irregulars and results in a strong increase of the net wind mass loading, while for high-mass galaxies it leads to the creation of bubbles and the development of a galactic fountain. Pressure forces within the disc reduce the kinetic energy in the wind above the disc by orders of magnitude. This paper is organized as follows. We present our recipe for galactic winds and its implementation in SPH simulations in section~\\ref{sec:recipe}. After describing our simulations of isolated disc galaxies in section~\\ref{sec:sims}, we present the results on the galaxies' morphology, star formation histories and large-scale winds in sections \\ref{sec:morphology} to \\ref{sec:winds}, respectively. In section \\ref{sec:resol} we present resolution tests which show that the effects of hydrodynamical drag are underestimated if the Jeans scale is unresolved. Finally, we summarize and discuss our conclusions in section~\\ref{sec:disc}. \\begin{table*} \\begin{center} \\caption{Simulation parameters: total mass, $M$; input mass loading, $\\eta$; input wind velocity, $v_{\\rm w}$; total number of particles, $N_{\\rm tot}$; total number of gas particles in the disc, $N_{\\rm disc}$; mass of baryonic particles, $m_{\\rm b}$; mass of dark matter particles, $m_{\\rm DM}$; gravitational softening of baryonic particles, $\\epsilon_{\\rm b}$; gravitational softening of dark matter particles, $\\epsilon_{\\rm DM}$; wind feedback included, (Wind); wind particles hydrodynamically decoupled, (Decoupled). Values different from the fiducial ones are shown in bold. \\label{tbl:params}} \\begin{tabular}{cccccccccccc} \\hline Simulation & $M_{\\rm halo}$ & $\\eta$ & $v_{\\rm w}$ & $N_{\\rm tot}$ & $N_{\\rm disc}$ & $m_{\\rm b}$ & $m_{\\rm DM}$ & $\\epsilon_b$ & $\\epsilon_{\\rm DM}$ & Wind & Decoupled\\\\ & $(\\Msolh)$ & & $(\\kms)$ & & & $(\\Msolh)$ & $(\\Msolh)$ & $(\\pch)$ & $(\\pch)$ & & \\\\ \\hline \\textit{m10} & $10^{10}$ & 2 & 600 & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^2$ & $2.4\\times 10^3$ & 10 & 17 & Y & N \\\\ \\textit{m10nowind} & $10^{10}$ & -- & -- & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^2$ & $2.4\\times 10^3$ & 10 & 17 & \\textbf{N} & -- \\\\ \\textit{m10$\\eta$1v848} & $10^{10}$ & \\textbf{1} & \\textbf{848} & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^2$ & $2.4\\times 10^3$ & 10 & 17 & Y & N \\\\ \\textit{m10$\\eta$4v424} & $10^{10}$ & \\textbf{4} & \\textbf{424} & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^2$ & $2.4\\times 10^3$ & 10 & 17 & Y & N \\\\ \\textit{m10dec} & $10^{10}$ & 2 & 600 & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^2$ & $2.4\\times 10^3$ & 10 & 17 & Y & \\textbf{Y} \\\\ \\hline \\textit{m10lr008} & $10^{10}$ & 2 & 600 & \\textbf{625$\\,$061} & \\textbf{29$\\,$411} & $\\mathbf{4.1\\times 10^3}$ & $\\mathbf{1.9\\times 10^4}$ & \\textbf{20} & \\textbf{34} & Y & N \\\\ \\textit{m10lr064} & $10^{10}$ & 2 & 600 & \\textbf{78$\\,$132} & \\textbf{3$\\,$676} & $\\mathbf{3.3\\times 10^4}$ & $\\mathbf{1.5\\times 10^5}$ & \\textbf{40} & \\textbf{68} & Y & N \\\\ \\textit{m10lr512} & $10^{10}$ & 2 & 600 & \\textbf{9$\\,$766} & \\textbf{459} & $\\mathbf{2.6\\times 10^5}$ & $\\mathbf{1.2\\times 10^6}$ & \\textbf{80} & \\textbf{136} & Y & N \\\\ \\hline \\textit{m12} & $10^{12}$ & 2 & 600 & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^4$ & $2.4\\times 10^5$ & 10 & 17 & Y & N \\\\ \\textit{m12nowind} & $10^{12}$ & -- & -- & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^4$ & $2.4\\times 10^5$ & 10 & 17 & \\textbf{N} & -- \\\\ \\textit{m12$\\eta$1v848} & $10^{12}$ & \\textbf{1} & \\textbf{848} & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^4$ & $2.4\\times 10^5$ & 10 & 17 & Y & N \\\\ \\textit{m12$\\eta$4v424} & $10^{12}$ & \\textbf{4} & \\textbf{424} & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^4$ & $2.4\\times 10^5$ & 10 & 17 & Y & N \\\\ \\textit{m12dec} & $10^{12}$ & 2 & 600 & 5000$\\,$494 & 235$\\,$294 & $5.1\\times 10^4$ & $2.4\\times 10^5$ & 10 & 17 & Y & \\textbf{Y}\\\\ \\hline \\textit{m12lr008} & $10^{12}$ & 2 & 600 & \\textbf{625$\\,$061} & \\textbf{29$\\,$411} & $\\mathbf{4.1\\times 10^5}$ & $\\mathbf{1.9\\times 10^6}$ & \\textbf{20} & \\textbf{34} & Y & N \\\\ \\textit{m12lr064} & $10^{12}$ & 2 & 600 & \\textbf{78$\\,$132} & \\textbf{3$\\,$676} & $\\mathbf{3.3\\times 10^6}$ & $\\mathbf{1.5\\times 10^7}$ & \\textbf{40} & \\textbf{68} & Y & N \\\\ \\textit{m12lr512} & $10^{12}$ & 2 & 600 & \\textbf{9$\\,$766} & \\textbf{459} & $\\mathbf{2.6\\times 10^7}$ & $\\mathbf{1.2\\times 10^8}$ & \\textbf{80} & \\textbf{136} & Y & N \\\\ \\hline \\end{tabular} \\end{center} \\end{table*} ", "conclusions": "\\label{sec:disc} Feedback from star formation is thought to play a key role in determining the observed properties of galaxies. It is therefore essential to include it in simulations of galaxy formation and evolution. Current simulations lack the resolution to resolve the energy conserving phase in the evolution of supernova remnants, causing any thermal energy input to be mostly radiated away before it has any significant hydrodynamical effect. Several types of sub-grid prescriptions for the generation of galactic winds have been proposed with the intent of overcoming numerical limitations. The most widely used methods use either thermal feedback combined with a temporary suppression of radiative cooling or kinetic feedback. In this work we introduced a sub-grid recipe to model feedback from massive stars in cosmological SPH simulations. The energy is distributed in kinetic form among the neighbours of recently formed stars. We implemented our prescription in the SPH code \\textsc{gadget} and tested it using high-resolution simulations of isolated disc galaxies of total mass $10^{10}$ and $10^{12}~\\Msolh$. Our kinetic feedback scheme strongly reduces the star formation rates and has a dramatic impact on the morphology of the galaxies. The disc of the dwarf galaxy becomes puffed up and punctuated with low-density bubbles, resembling the HI observations of nearby galaxies. A bipolar outflow develops naturally, with the largest velocities along the minor axis. The end result is a rather diffuse and irregular galaxy. Winds make the disc of the massive galaxy more stable, yielding more diffuse and less fragmented spiral arms. Its outer parts become more extended and also contain a large number of bubbles. While the wind is initially fastest along the minor axis, after a few hundred million years the velocities are higher at large opening angles because of strong infall along the minor axis. The edge-on gas density, temperature and pressure profiles remain, however, highly bi-conical. The infall is highly clumpy, consisting of cold gas clouds that have formed through thermal instabilities in the hot wind. The mean outflow velocity decreases with time, because gas that is blown out of the disc later must plough through gas that was blown out earlier, some of which has turned around and is falling back. At large radii the velocity increases with radius due to the fact that only gas with some minimum velocity can reach a given radius in the time since the initial starburst. The mean outflow velocities are much lower than the input values ($600~\\kms$ for our fiducial model), except at the largest radii reached by the wind. While the net mass outflow rate exceeds the input value ($ \\dot{M}_{\\rm w} / \\dot{M}_\\ast=2$ for our fiducial model) by about an order of magnitude for the dwarf galaxy, the reverse is true for the massive galaxy. Apparently, (ram) pressure forces in the disc enable the wind particles to drag large amounts of ISM out of the dwarf galaxy's disc, but are able to confine most wind particles to regions close to the disc in the case of the massive galaxy, whose ISM has both a higher density and pressure. We varied the input wind velocity from 424 to $848~\\kms$ while keeping the total kinetic energy per unit stellar mass formed constant by adjusting the input mass loading accordingly. For the dwarf galaxy the resulting star formation histories are all nearly identical and in the case of the massive galaxy only the lowest velocity run differed significantly. Thus, provided the input wind velocity is higher than some minimum value, which increases with the mass (and thus ISM pressure) of the galaxy, the results depend on the input kinetic energy but are insensitive to the amount of mass the energy is distributed over. While the star formation histories are relatively insensitive to numerical resolution, convergence of the predictions for the outflows requires resolving the Jeans scale. We contrasted our scheme with the sub-grid model of SH03 which has been widely used in the literature. In the SH03 prescription, the wind particles are selected stochastically from all the star-forming (i.e.\\ dense) particles in the simulation and are therefore non-local. These wind particles are subsequently decoupled from the hydrodynamics for 50~Myr (i.e.\\ 31~kpc if traveling at $600~\\kms$) or until their density has fallen below 10 percent of the threshold for star formation. In this work we have not tested the effects of non-locality because we expect them to be small for high-resolution simulations of individual galaxies such as those presented here. However, we do expect significant differences in cosmological simulations for which most galaxies will contain only a small number of star particles. In such galaxies the formation of star particles and the injection of kinetic energy will become essentially uncorrelated locally. To see this, note that for a galaxy that contains only a few star-forming particles, the typical time difference between the kicking of a wind particle and the creation of a star particle will be of order the gas consumption time scale divided by the input mass loading, or $\\sim 10^9/\\eta~\\yr$ at the threshold for star formation. This time scale greatly exceeds both the lifetime of massive stars and the simulation time step. The disconnection between the creation of a star particle and the injection of kinetic energy in its surroundings may have some undesirable consequences. For example, predictions for the chemical enrichment of the intergalactic medium may be affected because wind particles will contain an amount of heavy elements typical of particles in their host galaxies (which may often be zero) rather than abundances typical of the gas surrounding newly-formed stars. We carried out simulations in which the wind particles were temporarily decoupled from the hydrodynamics as in the SH03 prescription. The difference between the predictions for the fiducial and decoupled models is dramatic. Decoupled winds have almost no effect on the morphology of the disc. Compared with the fiducial model and HI observations of nearby galaxies, the dwarf galaxy, which has a low enough surface density to be stable, is much smoother while the massive galaxy, which is unstable without the injection of turbulence, is more clumpy but lacks low-density bubbles. While the winds in the fiducial model slightly increase the size of the gas disc, the decoupled winds continuously shrink the disc. While the coupled winds drive a large-scale bipolar outflow from the dwarf galaxy and a clumpy galactic fountain in the massive galaxy, the decoupled winds produce isotropic outflows in both cases. For the decoupled winds the outflow velocities (at least for $r < r_{\\rm vir}$) are constant in time and nearly independent of the galaxy mass. The mean, projected outflow velocity is about 70~percent of the input value, which greatly exceeds the outflow velocities observed in starbursting dwarf galaxies. The net mass outflow rates are in good agreement with the input values. The kinetic energy in the wind escaping the disc is orders of magnitude higher than in the case of the coupled winds. Compared with the fiducial model, the decoupled winds are less efficient at suppressing the SFR in the dwarf galaxy, but more efficient for the massive galaxy. The wind properties in the decoupled runs are insensitive to numerical resolution, even when the Jeans scale is completely unresolved. The lower the resolution of the runs with coupled winds, the more they resemble the case of decoupled winds. SH03 had several motivations for decoupling the winds hydrodynamically. They wanted to calibrate the sub-grid model with observations outside the disc, because they had no hope of resolving the structure of the ISM in their cosmological simulations, and they wanted a recipe that was insensitive to numerical resolution. Our tests clearly demonstrate that their method satisfies both these requirements. One may question, however, whether the observational uncertainties are not far too large to enable a calibration of the wind velocity and mass loading outside the disc. It is currently not even clear how the observed values depend on radius and gas phase. It is also questionable whether it is desirable for the predictions of hydrodynamical simulations to be insensitive to resolution if the simulations do not resolve the Jeans scale. Even if one is not interested in the internal structure of galaxies, which agrees less well with observations if the winds are decoupled, there are likely to be important differences in other types of predictions. The fact that hydrodynamic drag makes low-mass galaxies much more diffuse may, for example, greatly affect predictions for quasar absorption line observations and may also alleviate the angular momentum problem (i.e.\\ simulated discs are too small, probably due to excessive transfer of angular momentum to the dark matter halos). The fact that neglecting pressure forces on wind particles within the disc increases the kinetic energy of the escaping gas by very large factors, may have a large impact on predictions for the chemical enrichment of the intergalactic medium. However, we stress that our simulations lack the resolution and the physics necessary to predict the structure of the multiphase ISM and to model the small-scale effects that ultimately lead to the development of galactic winds. It is also important to note that while our artificial set-up of isolated, thin discs is useful for numerical experiments such as those presented here, it may exaggerate the differences between the coupled and decoupled winds. Galaxies in cosmological simulations are surrounded by gaseous halos and in the SH03 prescription wind particles will be re-coupled to the hydrodynamics before they leave the halos. Summarizing, our results suggest that (ram) pressure forces in the disc and the inner halo have a very strong impact on the structure of the ISM and the properties of galactic winds. Pressure forces exerted by expanding superbubbles puff up disc galaxies, give low-mass starbursting galaxies irregular morphologies and stabilize the discs of massive galaxies. The energy lost in this process strongly reduces the kinetic energy carried by the outflowing gas. Even if the first bubbles open up a channel in the disc through which the gas can be efficiently ejected, it will run into the gas which was blown out earlier and has been decelerated in the process. For massive galaxies the reduction in the kinetic energy results in the development of a galactic fountain. When the resolution is too low to resolve the Jeans scale, the effects of hydrodynamic drag on the galactic winds will be underestimated." }, "0801/0801.2400_arXiv.txt": { "abstract": "We analyze the absolute magnitude ($M_r$) and color ($u-r$) of low redshift ($z < 0.06$) galaxies in the Sloan Digital Sky Survey Data Release 6. Galaxies with nearly exponential profiles (Sloan parameter ${\\rm fracDeV} < 0.1$) all fall on the blue sequence of the color -- magnitude diagram; if, in addition, these exponential galaxies have $M_r < -19$, they show a dependence of $u-r$ color on apparent axis ratio $q$ expected for a dusty disk galaxy. By fitting luminosity functions for exponential galaxies with different values of $q$, we find that the dimming is well described by the relation $\\Delta M_r = 1.27 ( \\log q )^2$, rather than the $\\Delta M \\propto \\log q$ law that is frequently assumed. When the absolute magnitudes of bright exponential galaxies are corrected to their ``face-on'' value, $M_r^f = M_r - \\Delta M_r$, the average $u-r$ color is linearly dependent on $M_r^f$ for a given value of $q$. Nearly face-on exponential galaxies ($q > 0.9$) have a shallow dependence of mean $u-r$ color on $M_r^f$ (0.096 magnitudes redder for every magnitude brighter); by comparison, nearly edge-on exponential galaxies ($q < 0.3$) are 0.265 magnitudes redder for every magnitude brighter. When the dimming law $\\Delta M_r \\propto ( \\log q )^2$ is used to create an inclination-corrected sample of bright exponential galaxies, their apparent shapes are confirmed to be consistent with a distribution of mildly non-circular disks, with median short-to-long axis ratio $\\gamma \\approx 0.22$ and median disk ellipticity $\\epsilon \\approx 0.08$. ", "introduction": "\\label{sec-intro} Luminous galaxies ($M_r < -18$ or so) can be coarsely divided into two classes, conventionally labeled ``early-type'' and ``late-type''. Early-type galaxies have redder stellar populations and a scarcity of interstellar gas and dust. The majority of luminous early-type galaxies are elliptical galaxies, characterized by smooth isophotes and concentrated light profiles, well described by a \\citet{dV48} profile: $\\log I \\propto - r^{1/4}$. Highly luminous elliptical galaxies tend to be mildly triaxial ellipsoids (as opposed to perfectly oblate spheroids); their intrinsic short-to-long axis ratio is typically $c / a \\sim 0.7$ \\citep{ry92,vi05}. Late-type galaxies have blue stellar populations and relatively large amounts of interstellar gas and dust. The majority of luminous late-type galaxies are spiral galaxies, characterized by spiral structure within flattened disks. The disk light profile is generally well described by an exponential profile: $\\log I \\propto -r$. Luminous spiral galaxies are mildly elliptical (as opposed to perfectly circular) when seen face-on; their intrinsic short-to-long axis ratio is color-dependent, but at visible wavelengths is typically $c / a \\sim 0.25$ \\citep{bi81,gr85,la92,fa93,ry04,ry06}. In a color-magnitude (CM) diagram, if the color index is chosen correctly, the early-type and late-type galaxies manifest themselves as a ``red sequence'' and a ``blue sequence'', respectively. In the Sloan Digital Sky Survey (SDSS), the distribution of $u-r$ colors for low-redshift galaxies is bimodal; \\citet{st01} find an optimal color separator of $u-r = 2.22$, when color alone is used as a discriminator between early-type and late-type galaxies. Using a full CM diagram, the color separator is found to be dependent on $M_r$ \\citep{ba04}, with the color separator ranging from $u-r \\approx 2.3$ for galaxies with $M_r < -21$ to $u-r \\approx 1.8$ for galaxies with $M_r > -18$. However, a clean separation between early-type and late-type galaxies using color and absolute magnitude information alone is impossible; the red sequence and blue sequence overlap in a CM diagram. This overlap results partly from the fact that the color and apparent magnitude of spiral galaxies are inclination dependent. Since spiral galaxies are disk-shaped and contain dust, an edge-on spiral is both redder and fainter than the same spiral would be if seen face-on. Thus, as noted by \\citet{al02}, a sample chosen solely by the color criterion $u-r \\geq 2.22$ will contain dust-reddened edge-on spirals as well as intrinsically red ellipticals. The overlap between the red sequence and blue sequence would be reduced if we could perform an inclination correction on the colors and apparent magnitudes of spiral galaxies; that is, if we could convert observed apparent magnitudes into what they would be if the spiral galaxy were face-on. Since pivoting galaxies so that we can see them face-on is an impracticable task, we will take a statistical approach to finding the average dimming ($\\Delta M_r$) and reddening ($\\Delta (u-r)$) of a spiral galaxy as a function of its inclination $i$. In addition to allowing a cleaner separation between early-type and late-type galaxies in the CM diagram, a statistical correction for dimming and reddening has other practical uses. For instance, a flux-limited survey will undersample edge-on spirals with respect to face-on spirals; a galaxy that is just above the flux limit when it is face on would fall below the limit if it were edge on. With a knowledge of $\\Delta M_r$ as a function of inclination, it is possible to create an inclination-corrected flux-limited sample, by retaining only those galaxies that would still be above the flux limit if they were tilted to be seen edge on. Since the standard technique for finding the distribution of intrinsic axis ratios of galaxies assumes a random distribution of inclinations (see, for instance, \\citep{vi05} and references therein), such an inclination-corrected sample is essential for determining the true distribution of disk flattening. Without the inclination correction, the scarcity of edge-on galaxies would lead to an overestimate of the typical disk thickness. In section~\\ref{sec-data}, we describe how we select a sample of SDSS galaxies with exponential profiles; this gives us a population of disk-dominated spiral galaxies. The apparent axis ratio $q$ of the 25 mag arcsec$^{-2}$ isophote is chosen as our surrogate for the inclination of a galaxy. In section~\\ref{sec-analysis}, we examine the luminosity function of the SDSS exponential galaxies as a function of $q$. By shifting the luminosity functions until their high-luminosity cutoffs align, we can estimate the dimming $\\Delta M_r$ of the cutoff as a function of $q$. By brightening each galaxy by the $\\Delta M_r$ appropriate to its observed value of $q$, we can find its approximate ``face-on'' absolute magnitude $M_r^f$. We then provide linear fits for the mean $u-r$ color as a function of $M_r^f$ for different ranges of $q$. In section~\\ref{sec-shapes}, we create an inclination-corrected flux-limited sample, which we then use to find the distribution of intrinsic short-to-long axis ratios of the SDSS exponential galaxies, confirming that they are, in fact, disks, as our analysis assumed all along. Finally, in section~\\ref{sec-disc}, we provide a brief discussion of what the form of $\\Delta M_r$ as a function of $q$ and of $\\Delta (u-r)$ as a function of $q$ and $M_r^f$ implies for the properties of spiral galaxies and the dust they contain. ", "conclusions": "\\label{sec-disc} We have selected a population of galaxies from the Sloan Digital Sky Survey which are at low redshift ($z < 0.06$), which are relatively luminous ($M_r \\leq -19$), and which are well described by an exponential surface brightness profile (${\\rm fracDeV} < 0.1$). As we have shown, the properties of these galaxies are consistent with their being a population of slightly elliptical disks containing dust. The median dimensionless disk thickness for these galaxies in the $r$ band is $\\gamma \\approx 0.22$; the median disk ellipticity is $\\epsilon \\approx 0.08$. By fitting the luminosity function for galaxies with different apparent axis ratio $q$, we found that the apparent dimming $\\Delta M_r$ is not linearly proportional to $\\log q$, but instead is much better fitted by $\\Delta M_r \\propto ( \\log q )^2$. The dependence of dimming on inclination is a valuable clue to the dust properties within disk galaxies. If certain simplifying assumptions are made, the expected attenuation as a function of inclination can be computed for model galaxies. For instance, \\citet{fe99} assumed that dust had either the extinction curve found for Milky Way dust or for Small Magellanic Cloud dust \\citep{go97}. They assumed that disks were perfectly axisymmetric, with a horizontal scale length $r = 4 {\\rm\\,kpc}$ that was the same for both stars and dust. The scale height of the stars was assumed to be $z_\\star = 0.35$, but the scale height of the dust was allowed to vary. \\citet{fe99} found that dimming in the $B$ and $I$ bands were proportional to $\\log q$ only when the dust scale height was greater than that of the stars. However, observation of nearby edge-on disk galaxies \\citep{xi99} indicates that the dust scale height is about half the star scale height. \\citet{ro08}, using a Monte-Carlo radiative-transfer code to make calculations of internal extinction in dusty galaxies, found that a quadratic dependence of dimming on $\\log q$ provides a good fit for all plausible dust scale heights, scale lengths, and metallicity gradients. Since they were using hydrodynamic galaxy models with spiral structure, they were able to confirm that nonaxisymmetric structures such as spiral arms did not significantly affect the dependence of the total dimming $\\Delta M$ on the apparent axis ratio $q$. In our sample of exponential galaxies from the Sloan Digital Sky Survey, once the absolute magnitude of a galaxy is corrected for the inclination-dependent dimming, the mean $u-r$ color observed is linearly dependent on the corrected $M_r^f$. For nearly face-on galaxies, with $q > 0.9$, the dependence of $u-r$ on $M_r^f$ is relatively weak. We find $b = -0.096$; that is, less than 0.1 magnitude of reddening in $u-r$ for each magnitude brighter in $M_r^f$. For the edge-on galaxies, with $q < 0.3$, the dependence of $u-r$ on $M_r^f$ is much stronger, with $b = -0.265$. The mean color -- absolute magnitude dependence is a manifestation of the metallicity -- luminosity dependence. High-metallicity galaxies have both redder stellar populations and higher dust contents. For face-on exponential galaxies, the dust effects are minimized, and we see the effect of metallicity on stellar populations. For edge-on exponential galaxies, we see, in addition, the effect of metallicity on the dust content. Exponential galaxies with $M_r^f \\sim -21.5$ have a typical color $\\langle u - r \\rangle \\sim 1.8$ when seen face-on, placing them at the tip of the blue sequence in a color -- magnitude diagram. However, the same bright exponential galaxies when seen edge-on will have $M_r \\sim -21$ and $\\langle u -r \\rangle \\sim 2.5$, a degree of reddening that smuggles them into the red sequence, as usually defined. Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS website is \\url{http://www.sdss.org/}. The SDSS is managed by the Astrophysical Research Consortium (ARC) for the Participating Institutions. The Participating Institutions are the American Museum of Natural History, Astrophysical Institute Potsdam, University of Basel, University of Cambridge, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the Institute for Advanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, The Ohio State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University of Washington. \\newpage" }, "0801/0801.3226_arXiv.txt": { "abstract": "The dynamical interactions of planetary systems may be a clue to their formation histories. Therefore, the distribution of these interactions provides important constraints on models of planet formation. We focus on each system's apsidal motion and proximity to dynamical instability. Although only $\\sim$25 multiple planet systems have been discovered to date, our analyses in these terms have revealed several important features of planetary interactions. 1) Many systems interact such that they are near the boundary between stability and instability. 2) Planets tend to form such that at least one planet's eccentricity periodically drops to near zero. 3) Mean-motion resonant pairs would be unstable if not for the resonance. 4) Scattering of approximately equal mass planets is unlikely to produce the observed distribution of apsidal behavior. 5) Resonant interactions may be identified through calculating a system's proximity to instability, regardless of knowledge of angles such as mean longitude and longitude of periastron (\\eg GJ 317 b and c are probably in a 4:1 resonance). These properties of planetary systems have been identified through calculation of two parameters that describe the interaction. The apsidal interaction can be quantified by determining how close a planet is to an apsidal separatrix (a boundary between qualitatively different types of apsidal oscillations, \\eg libration or circulation of the major axes). This value can be calculated through short numerical integrations. The proximity to instability can be measured by comparing the observed orbital elements to an analytic boundary that describes a type of stability known as Hill stability. We have set up a website dedicated to presenting the most up-to-date information on dynamical interactions: http://www.lpl.arizona.edu/$\\sim$rory/research/xsp/dynamics. ", "introduction": "\\label{sec:intro} One of the most striking differences between known exoplanets and the giant planets of our Solar System involves the observed orbits: Observed exoplanets tend to have large eccentricities $e$, and small semi-major axes $a$, whereas the gas giants of the Solar System have small $e$ and large $a$. Recently, however, it has been shown that the dynamical interactions in many multiple planet systems (including the Solar System) show certain features in common (Barnes \\& Quinn 2004; Barnes \\& Greenberg 2006a [BG06a]; Barnes \\& Greenberg 2006c [BG06c]; Barnes \\& Greenberg 2007b [BG07b]). These shared traits suggest that the character of dynamical interactions (over $10^3$ -- $10^4$ years), rather than the present orbits, may be a more meaningful constraint on the origins of planetary systems (Barnes \\& Greenberg 2007a [BG07a], BG07b). Considerations of shared dynamical properties have even resulted in the first successful prediction of the mass and orbit of an extrasolar planet, HD 74156 d (predicted by Barnes \\& Raymond [2004] and Raymond \\& Barnes [2005]; found by Bean \\etal 2008). Now with over 200 extra-solar planets known, including $\\sim 25$ multi-planet systems, we can look at this population as a whole, with increasing confidence regarding which common characteristics may be more than statistical flukes. We have identified commonalities among planetary interactions that may be key constraints on the origins of planetary systems. We have identified parameters that quantify an interaction's proximity to boundaries between qualitatively different types of motion. These two types of boundaries are the ``apsidal separatrix'' and the dynamical stability boundary. The apsidal separatrix is the boundary between different types of apsidal oscillations, \\eg libration and circulation. The stability boundary separates regions in which all planets are bound to the host star from those in which at least one planet is liable to be ejected. These two parameters are fixed quantities that do not vary over time, but they constrain how the systems evolve. Systems tend to lie close to these two boundaries. In this chapter we review our derivation of $\\epsilon$, which quantifies proximity to an apsidal separatrix, in $\\S$ 2. Then we describe $\\beta$, which parameterizes a system's proximity to dynamical instability, in $\\S$ 3. In $\\S$ 4 we present the observed distributions of these quantities and use them to constrain formation models. Finally in $\\S$ 5 we draw our general conclusions. ", "conclusions": "\\label{sec:conclusions} We have described here new approaches for parameterizing the dynamical interactions of extrasolar planet systems. Many systems interact such that they are near the stability limit and the apsidal separatrix. Most mean-motion resonance interactions would be unstable if not for the resonance, a feature which may be used to identify resonant interactions (\\eg GJ 317 b and c). The distribution of $\\epsilon$ shows that the scattering of approximately equal mass planets is unlikely to produce the observed distribution of apsidal behavior. The distribution of $\\beta$ values shows that many systems formed near the limit of dynamical stability. These results demonstrate the benefits of the consideration of the dynamical interactions of multiple planet systems, which may constrain models of planet formation. For example, as we discussed above, the proximity of so many systems to an apsidal separatrix ($\\epsilon = 0$) suggests that ``rogue planets'' may have played a major role in scattering planets to higher eccentricities. The proximity of so many systems to the stability limit ($\\beta = 1$) suggests many systems form in densely packed configurations. Moreover, consideration of planetary systems in these terms has revealed that our Solar System shares dynamical traits with the known multiple planet systems. Perhaps constraining planet formation models through dynamical properties, rather than observed orbital elements, will lead to a universal model of planet formation. About half of planetary systems are multiple (Wright \\etal 2007), predictions of additional companions are being borne out (Barnes \\& Raymond 2004; Raymond \\& Barnes 2005; Bean \\etal 2008), and the current distribution of planet masses suggest there will be many planets with a mass equal to that of Saturn or less (Marcy \\etal 2005), \\ie below current detection limits. These three observations imply many multiple planet systems will be detected in the future. Hence characterizing extrasolar planet interactions will be a critical aspect of the study of planet formation for the foreseeable future." }, "0801/0801.1223_arXiv.txt": { "abstract": "CH$_4$ is proposed to be the starting point of a rich organic chemistry. Solid CH$_4$ abundances have previously been determined mostly toward high mass star forming regions. Spitzer/IRS now provides a unique opportunity to probe solid CH$_4$ toward low mass star forming regions as well. Infrared spectra from the Spitzer Space Telescope are presented to determine the solid CH$_4$ abundance toward a large sample of low mass young stellar objects. 25 out of 52 ice sources in the $c2d$ (cores to disks) legacy have an absorption feature at 7.7 $\\mu$m, attributed to the bending mode of solid CH$_4$. The solid CH$_4$ / H$_2$O abundances are 2-8\\%, except for three sources with abundances as high as 11--13\\%. These latter sources have relatively large uncertainties due to small total ice column densities. Toward sources with H$_2$O column densities above 2$\\times 10^{18}$ cm$^{-2}$, the CH$_4$ abundances (20 out of 25) are nearly constant at 4.7$\\pm$1.6\\%. Correlation plots with solid H$_2$O, CH$_3$OH, CO$_2$ and CO column densities and abundances relative to H$_2$O reveal a closer relationship of solid CH$_4$ with CO$_2$ and H$_2$O than with solid CO and CH$_3$OH. The inferred solid CH$_4$ abundances are consistent with models where CH$_4$ is formed through sequential hydrogenation of C on grain surfaces. Finally the equal or higher abundances toward low mass young stellar objects compared with high mass objects and the correlation studies support this formation pathway as well, but not the two competing theories: formation from CH$_3$OH and formation in gas phase with subsequent freeze-out. ", "introduction": "The presence and origin of complex organic molecules in protostellar regions and their possible incorporation in protoplanetary disks is an active topic of research. CH$_4$ is proposed to be a starting point of a rich chemistry, especially when UV photons are present \\citep{dartois05}. In particular CH$_4$ is believed to play a key role in the formation process of prebiotic molecules \\citep{markwick00}. CH$_4$ is less well studied in interstellar and circumstellar media compared to other small organic molecules because CH$_4$ has no permanent dipole moment and therefore cannot be observed by pure rotational transitions at radio wavelengths. Solid CH$_4$ was first detected through its bending mode at 7.67 $\\mu$m from the ground by \\citet{lacy91}, and with the Infrared Space Observatory Short Wavelength Spectrometer (ISO-SWS) by \\citet{boogert96} toward a few high-mass sources. Tentative claims have been made toward some other objects including low-mass protostars, but are inconclusive because of the low S/N ratio of these data \\citep{cernicharo00, gurtler02, alexander03}. Solid CH$_4$ has also been detected from the ground through its stretching mode at 3.3 $\\mu$m, but only toward the brightest high mass sources due to problems in removing the many atmospheric lines in this spectral region \\citep{boogert04}. Models predict CH$_4$ to form rapidly on cool grains through successive hydrogenation of atomic C; similarly H$_2$O is formed through hydrogenation of atomic O \\citep{vandehulst46, allen77, tielens82,brown88, hasegawa92,aikawa05}. Observations of CH$_4$ hence provide insight into the basic principles of grain surface chemistry. Compared to H$_2$O the observed gas- and solid-state CH$_4$ abundances are low; reported CH$_4$ abundances are typically a few percent with respect to H$_2$O \\citep{lacy91, boogert98}. This points to relatively low atomic C abundances at the time of CH$_4$ formation, with most C already locked up in CO as H readily reacts with C on surfaces \\citep{hiraoka98}. This is in agreement with the high CH$_3$OH abundances in several lines of sight, formed by hydrogenation of CO \\citep{dartois99,pontoppidan03}, and large CO$_2$ abundances, formed through oxidation of CO or hydrogenated CO. That these molecules are all formed through a similar process is corroborated by the profiles of solid CO$_2$ absorption bands, which usually show an intimate mixture of CO$_2$, CH$_3$OH and H$_2$O in interstellar ices \\citep{gerakines99, boogert00, knez05}. If CH$_4$ is formed efficiently through grain surface reactions as well, CH$_4$ should be similarly mixed with H$_2$O. Observations of solid CH$_4$ toward a few high mass young stellar objects (YSOs) show that the CH$_4$ absorption band profiles are broad and agree better with CH$_4$ in a hydrogen bonding ice, H$_2$O or CH$_3$OH, than with a pure CH$_4$ ice or CH$_4$ mixed with CO \\citep{boogert97}. This profile analysis does not, however, exclude CH$_4$ formation from photoprocessing of CH$_3$OH \\citep{allamandola88, gerakines96}. In addition, because of the small sample in previous studies, it is unclear if these broad profiles are a universal feature. Hence it cannot be excluded that CH$_4$ in some environments may form in the gas phase and subsequently freeze out. Because formation pathway efficiency depends on environment, another method for testing formation routes is through exploring the distribution of CH$_4$ toward a large sample of objects of different ages, luminosities and ice column densities. In addition, correlations, or lack thereof, with other ice constituents may provide important clues to how the molecule is formed. If CH$_4$ is formed through hydrogenation on grain surfaces in quiescent clouds, the CH$_4$ abundance with respect to H$_2$O should be fairly source independent since this mechanism mainly depends on the initial conditions of the cloud before the star forms, which seem to vary little between different star forming regions. Because of the generality of this mechanism, the CH$_4$ and H$_2$O, and possibly CO$_2$, column densities should correlate over a large range of different environments. This is also the prediction of several models where the solid CH$_4$/H$_2$O and CH$_4$/CO$_2$ ratios in dark clouds vary little both as a function of time \\citep{hasegawa92} and distance into a cloud that has collapsed \\citep{aikawa05}. In contrast, CO, which is formed in the gas phase and subsequently frozen out, is predicted to only correlate with CH$_4$ during certain time intervals. If CH$_4$ instead forms in the gas phase, solid CH$_4$ should be better correlated with solid CO than with solid H$_2$O and CO$_2$, since pure CH$_4$ freezes out and desorbs at similar temperatures to CO \\citep[Fraser et al. submitted to $A\\&A$,][]{collings04}. Finally, if CH$_4$ forms by UV photoprocessing of CH$_3$OH more CH$_4$ would also be expected to form toward sources with stronger UV-fields i.e. higher mass objects. The objective of this study is to determine the CH$_4$ abundances and distribution pattern toward a sample of low mass young stellar objects, varying in evolutionary stage and total ice column density. The distribution pattern and correlations with other ice constituents within the sample as well as comparison with high mass young stellar objects will be used to constrain the CH$_4$ formation mechanism. This study is based on spectra acquired with the Spitzer Infrared Spectrometer (IRS) as part of our legacy program `From molecular cores to protoplanetary disks' ($c2d$), which provides a large sample (41 sources) of infrared spectra of low mass star formation regions \\citep{evans03}. In addition 11 sources are added from the GTO program 2 for which ground based 3-5 $\\mu$m already exist \\citep{pontoppidan03}. Overviews of the H$_2$O, CO$_2$, CH$_3$OH and other ice species in these data are found in Boogert et al. (ApJ submitted, Paper I) and Pontoppidan et al. (ApJ submitted, Paper II). The detection of solid CH$_4$ toward one of the sources, HH46 IRS, was published by \\citet{boogert04b}. We have detected an absorption feature, which is attributed to solid CH$_4$, toward 25 out of 52 low mass ice sources found in this $c2d$ sample. ", "conclusions": "We present Spitzer-IRS spectra of the solid CH$_4$ feature at 7.7 $\\mu$m toward a large sample of low mass young stellar objects. Our conclusions are as follows: \\begin{itemize} \\item 25 out of 52 low mass young stellar objects show a solid CH$_4$ feature at 7.7 $\\mu$m. \\item The solid CH$_4$ abundance with respect to H$_2$O is centered at 5.8\\% with a standard deviation of 2.7\\% in the sources with CH$_4$ detections. In the sources without detections the average upper limit is 15\\%, which is not significant compared with the detections. \\item The sources (two Ophiuchus and one Serpens) with more than 10\\% CH$_4$ all have H$_2$O column densities below 2$\\times 10^{18}$ cm$^{-2}$. Due to the low total column densities, two of these three sources have uncertainities larger than 50\\%. Above 2$\\times 10^{18}$ cm$^{-2}$ the sources (20 out of 25) have a fairly constant CH$_4$ abundance of 4.7$\\pm$1.6\\%. \\item The 7.7$\\mu$m feature profiles are signficantly broader for all but one object than expected for pure solid CH$_4$ and toward most sources also broader than expected for CH$_4$ in H$_2$O dominated ices. Approximately 30\\% of the features have a blue wing, seen previously toward high mass YSOs and there attributed to solid SO$_2$ \\item The column densities of solid CH$_4$ and H$_2$O and CO$_2$ are clearly correlated, while CH$_4$ and CO and CH$_3$OH are only weakly correlated. \\item There is also no correlation between the CH$_4$ and CO abundances when both have been normalized to the H$_2$O abundance. \\item The Ophiuchus cloud has significantly higher CH$_4$ abundances compared to the rest of the sample, probably due to the low total column densities towards several of the sources. There are no significant differences between the remaining clouds. \\item The abundance variation is smaller for CH$_4$ compared to solid CH$_3$OH; CH$_4$ seems to belong to the class of molecules, also including H$_2$O and CO$_2$ that appear 'quiescent', i.e. their abundances are more or less constant, in contrast to highly variable ices like CH$_3$OH and OCN$^-$. If the Ophiuchus sources are included CH$_4$ is somewhere between the two classes. \\item Sample statistics and comparison with model predictions support CH$_4$ formation through hydrogenation of C on grain surfaces. \\end{itemize}" }, "0801/0801.4406_arXiv.txt": { "abstract": "We calculate the contribution of cosmic strings arising from a phase transition in the early universe, or cosmic superstrings arising from brane inflation, to the cosmic 21 cm power spectrum at redshifts $z\\geq 30$. Future experiments can exploit this effect to constrain the cosmic string tension $G\\mu$ and probe virtually the entire brane inflation model space allowed by current observations. Although current experiments with a collecting area of $\\sim 1$ $\\rm{km}^2$ will not provide any useful constraints, future experiments with a collecting area of $10^4-10^6$ $\\rm{km}^2$ covering the cleanest $10\\%$ of the sky can in principle constrain cosmic strings with tension $G\\mu \\gtrsim 10^{-10}-10^{-12}$ (superstring/phase transition mass scale $>10^{13}$ GeV). ", "introduction": " ", "conclusions": "" }, "0801/0801.1686_arXiv.txt": { "abstract": "We present and constrain a cosmological model which component is a pressureless fluid with bulk viscosity as an explanation for the present accelerated expansion of the universe. We study the particular model of a constant bulk viscosity coefficient $\\zeta_{{\\rm m}}$. The possible values of $\\zeta_{{\\rm m}}$ are constrained using the cosmological tests of SNe Ia Gold 2006 sample, the CMB shift parameter $R$ from the three-year WMAP observations, the Baryon Acoustic Oscillation (BAO) peak $A$ from the Sloan Digital Sky Survey (SDSS) and the Second Law of Thermodynamics (SLT). It was found that this model is in agreement with the SLT using only the SNe Ia test. However when the model is submitted to the three cosmological tests together (SNe+CMB+BAO) the results are: 1.- the model violates the SLT, 2.- predicts a value of $H_0 \\approx 53 \\; {\\rm km \\cdot sec^{-1} \\cdot Mpc^{-1}}$ for the Hubble constant, and 3.- we obtain a bad fit to data with a $\\chi^2_{{\\rm min}} \\approx 400$ ($\\chi^2_{{\\rm d.o.f.}} \\approx 2.2$). These results indicate that this model is ruled out by the observations. ", "introduction": " ", "conclusions": "" }, "0801/0801.4776_arXiv.txt": { "abstract": "We analyze the mid-infrared (MIR) spectra of ultraluminous infrared galaxies (ULIRGs) observed with the \\textit{Spitzer Space Telescope}'s Infrared Spectrograph. Dust emission dominates the MIR spectra of ULIRGs, and the reprocessed radiation that emerges is independent of the underlying heating spectrum. Instead, the resulting emission depends sensitively on the geometric distribution of the dust, which we diagnose with comparisons of numerical simulations of radiative transfer. Quantifying the silicate emission and absorption features that appear near 10 and 18\\um{} requires a reliable determination of the continuum, and we demonstrate that including a measurement of the continuum at intermediate wavelength (between the features) produces accurate results at all optical depths. With high-quality spectra, we successfully use the silicate features to constrain the dust chemistry. The observations of the ULIRGs and local sightlines require dust that has a relatively high 18/10\\um{} absorption ratio of the silicate features (around 0.5). Specifically, the cold dust of Ossenkopf et al. (1992) % is consistent with the observations, while other dust models are not. We use the silicate feature strengths to identify two families of ULIRGs, in which the dust distributions are fundamentally different. Optical spectral classifications are related to these families. In ULIRGs that harbor an active galactic nucleus, the spectrally broad lines are detected only when the nuclear surroundings are clumpy. In contrast, the sources of lower ionization optical spectra are deeply embedded in smooth distributions of optically thick dust. ", "introduction": "High infrared luminosity ($L_{IR} > 10^{12} L_\\sun$) characterizes ultraluminous infrared galaxies (ULIRGs). The underlying energy source may be accretion onto a supermassive black hole in an active galactic nucleus (AGN), intense bursts of star formation, or a combination of the two \\citep{San88a,Gen98}. Because the total luminosities are large, ULIRGs are critical sites to consider in obtaining a complete account of star formation, black hole growth, and the relationship between these two phenomena over cosmic time. In all cases, dust is responsible for reprocessing the intrinsic hard radiation to emerge at longer wavelengths. This dust reprocessing erases the spectral signatures that would reveal the nature of the original source, but the resulting spectra serve as probes of the dust itself. While most of the dust emission is continuum radiation, in the mid-infrared (MIR), two of its spectral features are observable, arising near 10 and 18\\um. These are attributed to silicates, especially amorphous analogs of pyroxenes and olivines \\citep[and references therein]{Dra03r}. Both features have been measured in stellar observations for decades \\citep*{Gil68,Woo69,Low70}. Observations of the stronger 10\\um{} absorption feature in starburst galaxies and AGNs have a similarly long history \\citep*[\\textit{e.g.},][]{Gil75,Rie75a,Rie75b,Kle76}, but only recently has the high sensitivity of instruments in space made measurement of the 18\\um{} feature in large numbers of galaxies feasible. Detecting silicate features in emission can identify the dusty medium as optically thin, and deep absorption characterizes optically thick obscuration. In detail, however, quantifying the optical depth requires a complete radiative transfer calculation because the same material produces both the continuum and line features. Thus, the apparent optical depth that silicate absorption strength yields is not the true optical depth along the line of sight, although it is frequently treated as such a direct measurement \\citep[\\textit{e.g.},][]{Shi06,Gal04,Mai01}. Compared with complete calculations, however, high signal-to-noise spectra of both of these features together effectively reveal the geometric distribution of the dust, including the total line-of-sight optical depth. Moreover, while the continuum emission depends only weakly on the dust chemistry, comparison of the two silicate features provides one of the few accessible diagnostics of the dust composition. Infrared studies that aim to deduce the nature of sources and surrounding dust geometry suffer from the degeneracy of the problem: a large range of different configurations produces very similar overall spectral energy distributions. Reaching conclusions about geometry, let alone dust chemistry, based on the spectral shape is difficult if not impossible \\citep[\\textit{e.g.}][]{Vin03}. However, extremely deep 10\\um{} silicate absorption proves to be a good discriminant of clumpy and smooth dust distributions, as we show in \\citet{Lev07}. Here we extend this analysis to include both the 10 and 18\\um{} silicate features. We apply detailed models to MIR spectral measurements of ULIRGs to probe the dust chemistry and geometric distribution of material around the galaxies' powerful nuclear sources. While the MIR spectra of ULIRGs have been studied previously \\citep[\\textit{e.g.},][]{Arm04,Arm06,Arm07,Des07,Far07,Mar07,Spo04,Spo06}, here we investigate both silicate features in detail. We consider the ULIRGs contained in the samples of \\citet{Spo07} and \\citet{Ima07}, which were all observed using the Infrared Spectrograph (IRS; \\citealt{Hou04}) on board the \\textit{Spitzer Space Telescope} \\citep{Wer04}. All the observations were obtained in low-resolution mode, which provides spectral resolution $R\\sim 100$ and complete coverage from 5--35\\um. The slits are up to 3.7 and 10.7\\arcsec{} wide in the short- and long-wavelength modules, respectively. Although these slits cover large areas of the galaxies (which have redshifts up to 0.93), the concentrated emission dominates the flux of the nuclear spectra we discuss. See \\citet{Spo07} for additional information about the data reduction. ", "conclusions": "The large luminosity of ULIRGs requires that these galaxies possess both an underlying source that is more energetic than star formation within ordinary galaxies and dust to reprocess the intrinsic radiation to long wavelengths. Observations in the IR offer the obvious advantage of directly detecting the light that emerges. Although they suffer the disadvantage that dust reprocessing erases the detailed signature of the nature of the energy source, as we have shown quantitatively, MIR spectroscopy does directly probe the dust itself. Analysis of the silicate features in particular provides useful diagnostics of the geometric distribution of dust and its optical properties, which ultimately depend on its mineralogy. Comparing observations of ULIRGs with several models of dust chemistry, we find that numerical simulations of radiative transfer employing the observationally-motivated OHMc dust properties match the data well, and this dust is required when deep absorption is measured. The key difference between this model and others that are frequently employed is the relatively high silicate absorption peak near 18\\um{} relative to that near 10\\um; this strength ratio is around 0.5 in the OHMc dust. These properties are physically-motivated, based on laboratory measurements of amorphous silicates. Including the effects of grain oxidation results in the higher ratio of 18/10\\um\\ cross section, which is characteristic of bronzite \\citep[and references therein]{Oss92}. This dust chemistry is not peculiar to the ULIRGs or to the nuclei of galaxies, but also fits observations within the Galaxy and Magellanic Clouds. Moreover, local observations where the silicates appear in emission rule out the OHMw and Draine dust at low optical depths. Quantifying the strength of the silicate features requires determining the underlying continuum accurately. We demonstrate that including measurements of the continuum at intermediate wavelengths between the features yields physically consistent results, and we recommend specific modifications to the idealized method when observational complications, such as photoionized emission or ice absorption bands, are present. We emphasize that because the same dust is responsible for both the feature and continuum, the apparent optical depth of silicate absorption does not reveal the line of sight optical depth, independent of the continuum-fitting procedure. We identify two distinct families in the ULIRG measurements, which is a consequence of fundamental differences in obscuring geometry. One class shows deep absorption ($S_{10}< -1$). These energetic sources must be deeply embedded in a continuous medium that is geometrically and optically thick ($\\tau_V > 100$). The other class exhibits weak silicate features, which may appear in absorption or emission. Either a geometrically thin slab or a clumpy dusty medium can account for these observations, but they do not represent the low optical depth realm of the continuous geometry. In addition, we find that these two ULIRG families of geometry are related to the optical identification of the galaxy nuclei. All members of the second (low-strength) class are AGNs, while ULIRGs having spectra characteristic of LINERs and \\ion{H}{2} regions systematically exhibit the deepest absorption. Among the deeply embedded sources, the MIR emission and optical lines always emanate from different regions. Optically-identified LINERs may be a heterogeneous class, but the consistent finding of \\ion{H}{2} ULIRGs only in the deeply embedded group leads us to speculate that two different regions dominate the IR and optical emission line signatures in all such optical \\ion{H}{2} ULIRGs. In spite of the general degeneracy of IR SEDs \\citep[\\textit{e.g.},][]{Vin03}, extreme absorption of the 10\\um{} feature alone proves to be a strong indicator of smooth (as opposed to clumpy) dust distributions \\citep{Lev07}. Here we find that the combination of 10 and 18\\um{} silicate features together constrains the geometry even more powerfully in a wider range of situations. Furthermore, the analysis tool of the feature-feature diagram we introduce here provides a strong diagnostic of the dust chemistry. The optical properties fix the absolute zero points and the slopes of the linear trends of silicate strength that variations of the model parameters can produce. However, because analysis based on this diagram is sensitive only to the feature strengths in the absorption cross section, it does not probe feature shape, peak wavelength, or properties such as mineral composition directly. We conclude that of the available models, the OHMc dust best describes cosmic silicate feature strengths, both in ULIRGs and local sources." }, "0801/0801.0964_arXiv.txt": { "abstract": "We present the work of an international team at the International Space Science Institute (ISSI) in Bern that worked together to review the current observational and theoretical status of the non-virialised X-ray emission components in clusters of galaxies. The subject is important for the study of large-scale hierarchical structure formation and to shed light on the \"missing baryon\" problem. The topics of the team work include thermal emission and absorption from the warm-hot intergalactic medium, non-thermal X-ray emission in clusters of galaxies, physical processes and chemical enrichment of this medium and clusters of galaxies, and the relationship between all these processes. One of the main goals of the team is to write and discuss a series of review papers on this subject. These reviews are intended as introductory text and reference for scientists wishing to work actively in this field. The team consists of sixteen experts in observations, theory and numerical simulations. ", "introduction": "Clusters of galaxies are the largest gravitationally bound structures in the Universe. Their baryonic composition is dominated by hot gas that is in quasi-hydrostatic equilibrium within the dark matter dominated gravitational potential well of the cluster. The hot gas is visible through spatially extended thermal X-ray emission, and it has been studied extensively both for assessing its physical properties and also as a tracer of the large-scale structure of the Universe. Clusters of galaxies are not isolated entities in the Universe: they are connected through a filamentary cosmic web. Theoretical predictions indicate the way this web is evolving. In the early Universe most of the gas in the web was relatively cool ($\\sim 10^4$~K) and visible through numerous absorption lines, designated as the so-called Ly$\\alpha$ forest. In the present Universe, however, about half of all the baryons are predicted to be in a warm phase ($10^5-10^7$~K), the Warm-Hot Intergalactic Medium (WHIM), with temperatures intermediate between the hot clusters and the cool absorbing gas causing the Ly$\\alpha$ forest. The X-ray spectra of clusters are dominated by the thermal emission from the hot gas, but in some cases there appears to be evidence for hard X-ray tails or soft X-ray excesses. Hard X-ray tails are difficult to detect, and one of the topics for the team is a discussion on the significance of this detection (yet contradictory) in existing and future space experiments. Various models have been proposed to produce these hard X-ray tails, and our team reviews these processes in the context of the observational constraints in clusters. While in some cases soft excesses in clusters can be explained as the low-energy extension of the non-thermal hard X-ray components mentioned above, there is evidence that a part may also be due to thermal emission from the WHIM. The signal seen near clusters then originates in the densest and hottest parts of the WHIM filaments, where the accelerating force of the clusters is highest and heating is strongest. A strong component of this emission is line radiation from highly ionised oxygen ions, and the role of this line emission and its observational evidence will be reviewed. WHIM filaments not only can be observed because of their continuum or line emission, but also through absorption lines if a sufficiently strong continuum background source is present. The evidence for absorption in both UV and X-ray high-resolution spectra is discussed. Future space missions will be well adapted to study these absorption lines in more detail. In particular in absorption lines the lower density parts between clusters become observable. In these low density regions of the WHIM not only collisional ionisation but also photo-ionisation is an important process. In general, the physics of the WHIM is challenging due to its complexity since there are many uncertain factors including the heating and cooling processes, the chemical enrichment, the role of supernova-driven bubbles or starburst winds, ram-pressure stripping, the role of shocks, magnetic fields, etc. More detailed (and sophisticated) hydrodynamical simulations with state-of-the-art spatial (and temporal) resolution are required in order to follow the impact of some (if not all) of these important processes. In particular chemical enrichment is an important process to consider as it leads to many observable predictions. We review the various physical processes relevant for the WHIM, the methods that are used to simulate this and the basic results from those models. ", "conclusions": "" }, "0801/0801.2759.txt": { "abstract": "We conducted a high resolution mid-infrared spectroscopic investigation using {\\it Spitzer} of 32 late-type (Sbc or later) galaxies that show no definitive signatures of Active Galactic Nuclei (AGN) in their optical spectra in order to search for low luminosity and/or embedded AGN. These observations reveal the presence of the high ionization [NeV] 14$\\mu$m and/or 24$\\mu$m line in 7 sources, providing strong evidence for AGNs in these galaxies. Taking into account the variable sensitivity of our observations, we find that the AGN detection rate based on mid-infrared diagnostics in optically normal late-type galaxies is $\\sim$ 30\\%, implying an AGN detection rate in late-type galaxies that is possibly 4 times larger than what optical spectroscopic observations alone suggest. We demonstrate using photoionization models with both an input AGN and an extreme EUV-bright starburst ionizing radiation field that the observed mid-infrared line ratios in our 7 AGN candidates cannot be replicated unless an AGN contribution, in some cases as little as 10\\% of the total galaxy luminosity, is included. These models show that when the fraction of the total luminosity due to the AGN is low, optical diagnostics are insensitive to the presence of the AGN. In this regime of parameter space, the mid-infrared diagnostics offer a powerful tool for uncovering AGN missed by optical spectroscopy. The AGN bolometric luminosities in our sample inferred using our [NeV] line luminosities range from $\\sim$ 3$\\times$10$^{41}$ ergs s$^{-1}$ to $\\sim$ 2$\\times$10$^{43}$ ergs s$^{-1}$. Assuming that the AGN is radiating at the Eddington limit, this range corresponds to a lower mass limit for the black hole that ranges from $\\sim$ 3$\\times$10$^3$M$_{\\odot}$ to as high as $\\sim$ 1.5$\\times$10$^5$M$_{\\odot}$. These lower mass limits however do not put a strain on the well-known relationship between the black hole mass and the host galaxy's stellar velocity dispersion established in predominantly early-type galaxies. Our findings add to the growing evidence that black holes do form and grow in low-bulge environments and that they are significantly more common than optical studies indicate. ", "introduction": "The vast majority of {\\it currently} known active galactic nuclei (AGN) in the local Universe reside in host galaxies with prominent bulges (e.g. Heckman 1980a; Keel 1983b; Terlevich, Melnick, \\& Moles 1987; Ho, Filippenko, \\& Sargent 1997; Kauffmann et al. 2003). This result, together with the finding that the black hole mass, M$_{\\rm BH}$, and the host galaxy's stellar velocity dispersion, $\\sigma$, is strongly correlated (Gebhardt et al. 2000; Ferrarese \\& Merritt 2000), has led to the general consensus that black hole formation and growth is fundamentally connected to the build-up of galaxy bulges. Indeed, it has been proposed that feedback from the AGN regulates the surrounding star formation in the host galaxy (e.g. Silk \\& Rees 1998; Kauffmann \\& Haehnelt 2000). However, an important outstanding question remains unresolved: {\\it is a bulge in general a necessary ingredient for a black hole to form and grow}? On the one hand, M33, the best studied nearby bulgeless galaxy, shows no evidence of a supermassive black hole (SBH), and the upper limit on the mass is significantly below that predicted by the M$_{\\rm BH}$-$\\sigma$ relation established in early-type galaxies (Gebhardt et al. 2001; Merritt et al. 2001). On the other hand, both NGC 4395 (Filippenko \\& Ho 2003) and POX52 (Barth et al. 2004) show no evidence for a bulge and yet do contain AGN. However, these two galaxies have remained isolated cases of bulgeless galaxies with accreting black holes, suggesting that they are anomalies. Indeed in the extensive Palomar optical spectroscopic survey of 486 nearby galaxies (Ho, Filippenko, \\& Sargent1997; henceforth H97), there are only 9 optically identified Seyferts with Hubble type of Sc or later. Only one galaxy of Hubble type of Scd or later in the entire survey is classified as a Seyfert (NGC 4395). Greene \\& Ho (2004, 2007) recently conducted an extensive search for broad-line AGN with intermediate mass black holes in the Fourth Data Release of the Sloan Digital Sky Survey (2004, 2007). Of the 8435 broad line AGN, they found only 174 (2\\%) such intermediate mass objects, indicating that they are extremely rare. Forty percent of these seem to be in late-type galaxies with colors consistent with morphological type Sab, although the Sloan images are of insufficient spatial resolution to extract precise morphological information. Optical observations thus clearly suggest that AGN in late-type galaxies are uncommon. However, determining if AGN reside in low bulge galaxies cannot be definitively answered using optical observations alone. The problem arises because a putative AGN in a galaxy with a minimal bulge is likely to be both energetically weak and deeply embedded in the center of a dusty late-type spiral. As a result, optical emission lines will be dominated by the emission from star formation regions, severely limiting the diagnostic power of optical surveys in determining the incidence of accreting black holes in low-bulge systems. In our previous work, we demonstrated the power of mid-IR spectroscopy in detecting the low-power AGN in the Sd galaxy NGC 3621 (Satyapal et al. 2007; henceforth S07). AGN show prominent high ionization fine structure line emission at mid-infrared wavelengths but starburst and normal galaxies are characterized by a lower ionization spectra characteristic of HII regions ionized by young stars (e.g. Genzel et al. 1998; Sturm et al. 2002; Satyapal et al. 2004). In particular, the [NeV] 14 $\\mu$m (ionization potential 97 eV) line is not generally produced in HII regions surrounding young stars, the dominant energy source in starburst galaxies, since even hot massive stars emit very few photons with energy sufficient for the production of Ne4+. The detection of this line in any galaxy provides strong evidence for an AGN. The goal of this paper is to answer the question: are AGN in late-type galaxies more common than previously thought? If so, this would revise our understanding of the environments in which supermassive black holes (SMBHs) can form and grow and perhaps shed light on the nature of the M$_{\\rm BH}$-$\\sigma$ relation in low-bulge systems. Toward this end, we conducted an archival investigation of 32 late-type galaxies observed by the {\\it Spitzer} high resolution spectrograph that show no definitive optical signatures of AGN to search for low-power and/or deeply embedded AGN. This paper is structured as follows. In Section 2, we summarize the properties of the {\\it Spitzer} archival sample presented in this paper. In Section 3, we summarize the observational details and data analysis procedure, followed by a description of our results in Section 4. In Section 5, we discuss the origin of the [NeV] emission , with each galaxy discussed individually, followed by a discussion of the implications of our discoveries in Section 6. A summary of our major conclusions is given in Section 7. ", "conclusions": "We conducted a mid-infrared spectroscopic investigation of 32 late-type (Hubble type of Sbc or later) galaxies showing no definitive signatures of AGN in their optical spectra in order to search for low luminosity and/or embedded AGN. The primary goal of our study was determine if AGN in low-bulge environments are more common than once thought. Our high resolution {\\it Spitzer} spectroscopic observations reveal that the answer to this question is {\\it yes}. Our main results are summarized below: \\begin{enumerate} \\item We detected the high ionization [NeV] 14.3$\\mu$m and/or 24.3$\\mu$m lines in 7 late-type galaxies, providing strong evidence for AGNs in these galaxies. \\item We detected the high excitation [OIV] 25.9$\\mu$m and [NeIII] 15.5$\\mu$m lines in 5 out of the 7 of the galaxies with [NeV] emission. Although these lines can be excited in star forming regions, our mapping observations (when available) suggest that the emission is centrally concentrated and likely to be dominated by the AGN. \\item Taking into account the range of sensitivities of our observations, our work suggests that the AGN detection rate based on mid-infrared diagnostics in late-type optically normal galaxies can be as much or more than $\\sim$ 30\\%. This detection rate implies that the overall fraction of late-type galaxies hosting AGN is possibly more than 4 times larger than what optical spectroscopic observations alone suggest. \\item Several of the galaxies with [NeV] detections have optical emission line ratios in the extreme ``starburst range'' indicating that there is absolutely no hint of an AGN based on their optical spectra. Three out of the 7 galaxies are classified based on their optical line ratios as ``transition objects'' but none show broad permitted optical lines. \\item Amongst the 7 AGN candidates in our sample, 3 are Sbc, 3 are Sc, and 1 is of Hubble type Scd. Since there are only 3 galaxies of Hubble type Sd, our limited sample size precludes us from making any definitive conclusions on the incidence of AGN in completely bulgeless galaxies. The lowest central stellar velocity dispersion amongst the galaxies with published measurements is 40 km/s. \\item We demonstrate using photoionization models with both an input AGN and an extreme EUV-bright starburst ionizing radiation field that the observed mid-infrared line ratios in our 7 AGN candidates cannot be replicated unless an AGN contribution is included. These models show that when the fraction of the total luminosity due to the AGN is low, the optical diagnostics are insensitive to the presence of the AGN. In this regime of parameter space, the mid-infrared diagnostics offer a powerful tool in uncovering AGN missed by optical spectroscopy. \\item Three of the galaxies, NGC 3556, NGC 3367, and NGC 4536, appear to be dominated by star-formation. NGC 3556 could have as little as 10\\% of its luminosity coming from the AGN. NGC 4321, NGC 5055, NGC 3938, and NGC 4414 likely have a more dominant contribution to their luminosity from the AGN in addition to having some contribution to their emission line fluxes from shock-excited gas. All of the galaxies that H97 classifies as HII galaxies are well characterized by a low AGN contribution, while all the transition objects seem to require some shock component. \\item The AGN bolometric luminosities inferred using our [NeV] line luminosities range from $\\sim$ 3$\\times$10$^{41}$ ergs s$^{-1}$ to $\\sim$ 2$\\times$10$^{43}$ ergs s$^{-1}$, with a median value of $\\sim$ 9$\\times$10$^{41}$ ergs s$^{-1}$. Assuming that the AGN is radiating at the Eddington limit, this corresponds a lower mass limit for the black hole that ranges from $\\sim$ 3$\\times$10$^3$M$_{\\odot}$ to as high as $\\sim$ 1.5$\\times$10$^5$M$_{\\odot}$. These lower mass limits however do not put a strain on the well-known relationship between the black hole mass and the host galaxy\u0092s stellar velocity dispersion established in predominantly early-type galaxies. \\end{enumerate} The {\\it Spitzer} spectroscopic study presented here demonstrates that black holes do form and grow in low-bulge environments and that they are significantly more common than optical studies indicate. In order to truly determine how common SBHs and AGN activity are in {\\it completely} bulgeless galaxies, a more extensive study with {\\it Spitzer} is crucial." }, "0801/0801.0013_arXiv.txt": { "abstract": "The Vector Spectromagnetograph (VSM) instrument has been recording photospheric and chromospheric magnetograms daily since August 2003. Full-disk photospheric vector magnetograms are observed at least weekly and, since November 2006, area-scans of active regions daily. Quick-look vector magnetic images, plus X3D and FITS formated files, are now publicly available daily. In the near future, Milne-Eddington inversion parameter data will also be available and a typical observing day will include three full-disk photospheric vector magnetograms. Besides full-disk observations, the VSM is capable of high temporal cadence area-scans of both the photosphere and chromosphere. Carrington rotation and daily synoptic maps are also available from the photospheric magnetograms and coronal hole estimate images. ", "introduction": "The VSM currently operates at Kitt Peak, Arizona and is part of the Synoptic Optical Long-term Investigations of the Sun (SOLIS) project \\citep{Keller03}. The VSM provides a unique record of solar full-disk vector magnetograms along with high-sensitivity photospheric and chromospheric longitudinal magnetograms \\citep{Henney06}. The VSM began recording full-disk vector magnetograms in August 2003 at a temporary site in Tucson, Arizona. In April 2004, the VSM was relocated to Kitt Peak and resumed operations in May 2004. The VSM records the full Stokes profiles of the Fe I 630.15 and 630.25 nm lines with a current spatial and spectral sampling of 1.125 arcsec and 2.71 pm respectively. The 50-cm aperture VSM utilizes a Ritchey-Chr\\' etien optical design, where full-disk images are constructed from 2048 individual steps in declination of the projected solar image on the entrance slit (hereafter referred to as scan-lines). Observations with fewer scan-lines are referred to as area-scans and are typically recorded for active region areas. For example, 88 area-scans were recorded in the month of October 2003 during an exceptionally active flare period. The VSM is available for user-requested programs and currently supports ongoing solar missions (e.g. MDI, STEREO, Hinode/SOT) and will support forthcoming missions (e.g. SDO/HMI). For example, VSM observations provide temporal and spatial context for the SOT observations. Though SOT/SP has exceptional spatial resolution, the VSM observes any solar region eight times faster than SOT/SP and ten times larger in one spatial direction. The VSM also provides full-disk coverage not available with SOT. In addition, the future SDO/HMI instrument will have high temporal cadence and spatial resolution, but will have very limited spectral resolution (with six filter images) that the VSM will complement with fully resolved Stokes profiles. \\begin{figure*}[!ht] \\plotone{henney_fig1_v1.ps} \\caption{An example VSM vector quick-look image of NOAA active region 10921, observed on November 3, 2006. The gray-scale illustrates the line-of-sight (LOS) magnetic flux and the arrows indicate the transverse field strength (the length scale is shown in the lower left-hand corner) and direction.} \\label{QL-AR} \\end{figure*} \\subsection{VSM Upgrades} Starting in 2004, the VSM suffered from a slow degradation of the polarization modulator needed to measure full Stokes profiles. This complicated VSM vector calibrations and slowed efforts to release the vector magnetic data. In February of 2006, the vector modulator was replaced along with two modulators for the line-of-sight component of the magnetic field in the solar photosphere (630.2 nm) and the chromosphere (854.2 nm). During this maintenance period, a broken part was also replaced to reduce grating flexure. Though the new modulators are performing well, e.g. the strength of the 854.2 nm signal has significantly increased, the vector modulator has a polarization fringe pattern that is temperature dependent. The fringe pattern has recently been well parameterized to minimize the number of terms required for a good fit. The fringe fitting algorithm is expected to be incorporated in the SOLIS processing pipeline in early 2008. Since 2003, the VSM has incorporated two interim CMOS hybrid cameras made by Rockwell Scientific (now Teledyne Scientific \\& Imaging, LLC). During the first year of operation, several unexpected camera signals were revealed. Two of the pronounced artifacts, a variable dark level and signal cross-talk at the 0.5\\% level, required significant modification to the data reduction pipeline. The cameras were also found to have a 25\\% residual signal from previous frames. These interim cameras are scheduled to be replaced early in 2008 with cameras produced by Sarnoff that better match the spatial scale, sensitivity, low readout noise and desired frame rate for the VSM. ", "conclusions": "" }, "0801/0801.2691_arXiv.txt": { "abstract": "{The observed clumpy structures in debris disks are commonly interpreted as particles trapped in mean-motion resonances with an unseen exo-planet. Populating the resonances requires a migrating process of either the particles (spiraling inward due to drag forces) or the planet (moving outward). Because the drag time-scale in resolved debris disks is generally long compared to the collisional time-scale, the planet migration scenario might be more likely, but this model has so far only been investigated for planets on circular orbits.} {We present a thorough study of the impact of a migrating planet on a planetesimal disk, by exploring a broad range of masses and eccentricities for the planet. We discuss the sensitivity of the structures generated in debris disks to the basic planet parameters.} {We perform many N-body numerical simulations, using the symplectic integrator SWIFT, taking into account the gravitational influence of the star and the planet on massless test particles. A constant migration rate is assumed for the planet.} {The effect of planetary migration on the trapping of particles in mean motion resonances is found to be very sensitive to the initial eccentricity of the planet and of the planetesimals. A planetary eccentricity as low as $0.05$ is enough to smear out all the resonant structures, except for the most massive planets. The planetesimals also initially have to be on orbits with a mean eccentricity of less than than $0.1$ in order to keep the resonant clumps visible.} {This numerical work extends previous analytical studies and provides a collection of disk images that may help in interpreting the observations of structures in debris disks. Overall, it shows that stringent conditions must be fulfilled to obtain observable resonant structures in debris disks. Theoretical models of the origin of planetary migration will therefore have to explain how planetary systems remain in a suitable configuration to reproduce the observed structures.} ", "introduction": "Since the first direct imaging of a debris disk around \\object{$\\beta$ Pictoris} by \\citet{1984Sci...226.1421S}, a dozen other optically thin dust disks have been spatially resolved around nearby main-sequence stars showing an infrared excess \\citep[][and references there in]{2007ApJ...661L..85K,2006ApJ...650..414S}. The images often reveal asymmetric structures and clumps, interpreted as the signature of gravitational perturbations. A planet immersed in a debris disk usually produces structures such as a gap along its orbit, by ejecting particles during close encounters, or density waves (e.g. a one-arm spiral), by modifying the precession rate of the dust particles \\citep{2005A&A...440..937W}. However, such structures cannot explain the observations of clumpy, non-axisymmetric disks \\citep{2004ASPC..321..305A,2007prpl.conf..573M}, and resonant mechanisms with unseen planets have been proposed to account for the observed asymmetries \\citep{2000ApJ...537L.147O,2002ApJ...578L.149Q,2003ApJ...588.1110K,2003ApJ...598.1321W}. A particle belongs to a mean motion resonance (MMR) when the particle to planet period ratio is a rational number, $m:n$ with $m$ and $n$ integers. An MMR is located at a semi-major axis $a$ given by $a/a_p=(m/n)^{2/3}$, where $a_p$ is the planet semi-major axis. In the Solar System, for example, about $15\\%$ of the known Kuiper Belt objects, including Pluto, are trapped in the $3$:$2$ resonance with Neptune \\citep{2007prpl.conf..895C}. The interesting property of MMRs for modeling asymmetric disks is that, as explained for example in \\citet{2000ssd..book.....M}, resonant objects are not uniformly distributed in azimuth around a star: rather they gather at specific longitudes relative to the perturbing planet and subsequently form clumps. This arises from properties specific to MMRs as a given particle trapped in a MMR with a planet undergoes conjunctions with the planet at specific locations along its orbit. The particles tend to gather around the most stable orbital configurations that ensure that the conjunctions occur at the maximum relative distance. The clumps, which are the result of the collective effects of resonant particles, generally corotate with the planet \\citep{2003ApJ...588.1110K}, while each of these resonant bodies has a different period from that of the planet (except for $1$:$1$ resonant planetesimals): hence the motion of these density waves differs from the orbital motion of the resonant particles. However, MMRs are very thin radial structures that usually trap a small number of particles in a given disk. Therefore, any structure due to MMRs has a high chance of being totally hidden by the emission of the non-resonant particles, as illustrated in Fig. \\ref{withoutMigration}. For clumps due to MMRs to be observed, the population of resonant particles must be significantly enhanced by an additional physical process. Two mechanisms can account for this: Poynting-Robertson (P-R) drag and planet migration. Dust particles that are too large to be ejected from the system by radiation pressure can spiral inward into the star due to P-R drag and to some other minor forces like stellar wind drag \\citep[e.g.][]{2006A&A...455..987A}. In the course of its inward migration, a dust particle can be trapped into exterior MMRs with a planet, hence increasing the contrast of their asymmetric patterns \\citep{2003ApJ...588.1110K,2005ApJ...625..398D}. Planet migration, on the other hand, involves particles of all sizes, except those ejected by radiation pressure. Many particles can be trapped in MMRs by a planet migrating outward in the disk. Each non-resonant particle crossing an MMR has a chance trapped and subsequently migrating, following the resonance \\citep{2003ApJ...598.1321W}. Several authors have studied either the effect of P-R drag, or of planet migration, on disk structures, using different methods (analytic, semi-analytic or numerical) and various planet parameters (mass and orbital eccentricity). A summary of the main previous studies is provided in Table \\ref{previousWorks}. The P-R drag scenario has been extensively studied for a wide range of parameters, while the migrating planet scenario has been investigated only for a planet on a circular orbit by \\citet{2003ApJ...598.1321W}. It is thus important to better characterize the latter scenario in order to distinguish which of the two dominates the morphology of debris disks. Moreover, a number of studies \\citep{1996Icar..123..168L,2005A&A...433.1007W,2007A&A...462..199K} have shown that collisions may prevent MMRs from being populated by P-R drag since the collision timescale in massive debris disks might be much shorter than the P-R drag migration timescale. Therefore, we propose in this paper to extensively study the planet migration scenario, using numerical modeling, by generating a synthetic catalog similar to what has been done for the P-R drag scenario using analytical \\citep{2003ApJ...588.1110K} or numerical studies \\citep{2005ApJ...625..398D}. We extend the pioneering work done by \\citet{2003ApJ...598.1321W} in studying the influence of the planet eccentricity on the visibility of the resonant patterns. In Section \\ref{lowEccOrb}, we discuss the case of a planet migrating on a circular or low-eccentricity orbit. In Section \\ref{highEccOrb}, we extend this study to planets on orbits with eccentricities up to $0.7$. This study is generalized to various migration rates and disk initial states in Section \\ref{Generalization}, and compared to previous studies in Section \\ref{Comparison}. The limitations of our approach are discussed in Section \\ref{Limitation}. \\begin{table*} \\centering \\caption{\\label{previousWorks}Summary of recent papers on particles trapping in MMRs with a planet.} \\begin{tabularx}{\\textwidth}{c c c c c X} Authors & Method & \\multicolumn{2}{c}{Planet parameters} & Migration & Notes \\\\ && Mass ratio$^a$ & eccentricity \\\\ \\hline\\hline \\\\ \\citet{2003ApJ...598.1321W} & semi-analytic & $0.0003$ to $3$& $0$ & planet & Forced migration\\\\ \\citet{2006ApJ...639.1153W} & semi-analytic &&& none & This work extends the previous one to smaller particles sensitive to radiation pressure.\\\\ \\citet{2003ApJ...588.1110K} & analytic & $0.005$ to $15$ & $0$ to $0.6$ & particles & Only resonant particles trapped during migration due to P-R drag. \\\\ \\citet{2005ApJ...625..398D} & numerical & $0.01$ to $3$ & $0$ to $0.7$& particles & Particles migrate due to P-R drag and solar wind. \\\\ \\citet{2002AJ....124.2305M} & numerical &$0.05$&$0$& particles & Study of the Kuiper Belt. \\\\ This paper & numerical & $0.001$ to $3$ & $0$ to $0.7$ & planet & Only planetesimals disk, forced migration \\\\ \\hline \\end{tabularx} \\begin{list}{}{} \\item[$^a$]Planetary mass in Jovian mass divided by stellar mass in solar mass. \\end{list} \\end{table*} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics[angle=-90]{figures/withoutMigration.eps}} \\caption{\\label{withoutMigration} Example of a planetesimal disk without outward planet migration, nor inward P-R drag migration of the test particles, according to our numerical simulations as explained in Section \\ref{model}. The star and planet locations, projected onto the orbital plane of the planet, are represented by large red points, and the planet orbit by a thin green line. The initial planetesimal disk consists of $50\\,000$ planetesimals distributed between 40 and 75 AU, with the surface density distribution proportional to $r^{-1}$. Although some planetesimals are trapped in MMRs with the planet, they are not sufficiently numerous to generate spatial structures (besides the $1$:$1$ MMR). \\thanks{See the electronic edition of the Journal for a color version of this figure.}} \\end{figure} ", "conclusions": "We have studied the problem of the presence of observable structures in planetesimal disks due to mean motion resonance with an unseen planet migrating outward in the disk. Using numerical simulations, we have explored a large range of parameters for the planet (mass and orbital eccentricity) and the disk (initial distribution of planetesimal eccentricities). In the case of a planet on a circular orbit migrating inside a dynamically cold disk, our results are in agreement with previous analytical studies. In the cases not already addressed, namely planets on eccentric orbits or dynamically warm disks, we have found that the observability of resonant structures demands very specific orbital configurations. The clumps produced by MMRs with a planet on a circular orbit are smoothed in the case of a planet on an even moderately eccentric orbit. An eccentricity as low as $0.05$ is enough to smooth all the resonant structures, except for the most massive planets. These results indicate that although trapping planetesimals in MMRs is an efficient mechanism to generate clumpy disks, stringent conditions must be fulfilled for this scenario to occur. Theoretical modeling of the origin of the planetary migration therefore will have to explain how planetary systems can remain under these conditions. Moreover, we only consider a planet migrating at a constant rate. A more realistic model with a variable, stochastic migration rate can reduce the population of resonances and thus their observability. A better model of planet migration thus should be developed in future studies. \\begin{table*} \\centering \\caption{\\label{normal} Summary of results for all simulations done in the present study with an initially unexcited disk (initial planetesimal eccentricities are equal to zero) and a standard migration rate of $0.5$ AU Myr$^{-1}$. For each simulation, we list the migration rate, the planet mass and eccentricity, the fraction of surviving planetesimals at the end of the simulation ($40$ Myr) and the resulting disk shape, following the convention of Fig. \\ref{resume}.} \\begin{tabular}{ccccc||ccccc} Mig. rate$^a$ & Mass$^b$ & Ecc. & Surv. planetesimals$^c$ & Disk shape$^d$& Mig. rate$^a$ & Mass$^b$ & Ecc. & Surv. planetesimals$^c$ & Disk shape$^d$\\\\ \\hline 0.5 & 0.0035 &0.0&$100\\%$&None&0.5 & 0.05 &0.0&$100\\%$&I\\\\ & &0.01&$100\\%$&None& & &0.01&$100\\%$&I\\\\ & &0.05&$100\\%$&None& & &0.05&$100\\%$&II\\\\ & &0.1&$100\\%$&None& & &0.1&$100\\%$&None\\\\ & &0.2&$100\\%$&None& & &0.2&$50\\%$&None\\\\ & &0.3&$100\\%$&None& & &0.3&$10\\%$&None\\\\ & &0.4&$35\\%$&III& & &0.4&$5\\%$&None\\\\ & &0.5&$10\\%$&III& & &0.5&$10\\%$&None\\\\ & &0.6&$5\\%$&III& & &0.6&$5\\%$&None\\\\ & &0.7&$5\\%$&III& & &0.7&$5\\%$&None\\\\ 0.5 & 0.33 &0.0&$75\\%$&I&0.5 & 1 &0.0&$70\\%$&I\\\\ & &0.01&$70\\%$&I& & &0.01&$70\\%$&I\\\\ & &0.05&$85\\%$&II& & &0.05&$55\\%$&II\\\\ & &0.1&$75\\%$&None& & &0.1&$25\\%$&II\\\\ & &0.2&$25\\%$&None& & &0.2&$10\\%$&None\\\\ & &0.3&$10\\%$&None& & &0.3&$10\\%$&None\\\\ & &0.4&$10\\%$&None& & &0.4&$10\\%$&None\\\\ & &0.5&$10\\%$&None& & &0.5&$5\\%$&None\\\\ & &0.6&$5\\%$&None& & &0.6&$5\\%$&None\\\\ & &0.7&$5\\%$&None& & &0.7&$0\\%$&None\\\\ 0.5 & 2 &0.0&$65\\%$&I&0.5 & 3 &0.0&$55\\%$&I\\\\ & &0.01&$60\\%$&I& & &0.01&$50\\%$&I\\\\ & &0.05&$30\\%$&I& & &0.05&$20\\%$&I\\\\ & &0.1&$15\\%$&II& & &0.1&$10\\%$&I\\\\ & &0.2&$5\\%$&None& & &0.2&$0\\%$&None\\\\ & &0.3&$5\\%$&None& & &0.3&$0\\%$&None\\\\ & &0.4&$5\\%$&None& & &0.4&$0\\%$&None\\\\ & &0.5&$5\\%$&None& & &0.5&$0\\%$&None\\\\ & &0.6&$0\\%$&None& & &0.6&$0\\%$&None\\\\ & &0.7&$0\\%$&None& & &0.7&$0\\%$&None\\\\ \\hline \\multicolumn{10}{l}{$^a$In AU Myr$^{-1}$.}\\\\ \\multicolumn{10}{l}{$^b$In Jovian mass.}\\\\ \\multicolumn{10}{l}{$^c$Fraction of surviving planetesimals at the end of the simulation, i.e. $40$ Myr.}\\\\ \\multicolumn{10}{l}{$^d$As in Fig. \\ref{resume}.}\\\\ \\end{tabular} \\end{table*} \\begin{table*} \\centering \\caption{\\label{fastslow} Same as Table \\ref{normal}, but for different migration rates.} \\begin{tabular}{ccccc||ccccc} Mig. rate$^a$ & Mass$^b$ & Ecc. & Surv. planetesimals$^c$ & Disk shape$^d$& Mig. rate$^a$ & Mass$^b$ & Ecc. & Surv. planetesimals$^c$ & Disk shape$^d$\\\\ \\hline 5 & 0.035 &0.0&$100\\%$&None&5 & 0.05 &0.0&$100\\%$&I\\\\ & &0.01&$100\\%$&None& & &0.01&$100\\%$&II\\\\ & &0.05&$100\\%$&None& & &0.05&$100\\%$&II\\\\ & &0.1&$100\\%$&None& & &0.1&$100\\%$&None\\\\ & &0.2&$100\\%$&None& & &0.2&$100\\%$&None\\\\ 5 & 0.33 &0.0&$100\\%$&I&5 & 1 &0.0&$100\\%$&I\\\\ & &0.01&$100\\%$&I& & &0.01&$100\\%$&I\\\\ & &0.05&$100\\%$&II& & &0.05&$100\\%$&II\\\\ & &0.1&$100\\%$&II& & &0.1&$95\\%$&II\\\\ & &0.2&$95\\%$&None& & &0.2&$75\\%$&None\\\\ 5 & 2 &0.0&$85\\%$&I&5 & 3 &0.0&$75\\%$&I\\\\ & &0.01&$85\\%$&I& & &0.01&$75\\%$&I\\\\ & &0.05&$80\\%$&I& & &0.05&$60\\%$&I\\\\ & &0.1&$65\\%$&II& & &0.1&$40\\%$&I\\\\ & &0.2&$50\\%$&None& & &0.2&$25\\%$&None\\\\ 0.05 & 0.035 &0.0&$100\\%$&II&0.05 & 0.05 &0.0&$90\\%$&I\\\\ & &0.01&$100\\%$&II& & &0.01&$85\\%$&I\\\\ & &0.05&$100\\%$&None& & &0.05&$85\\%$&II\\\\ & &0.1&$100\\%$&None& & &0.1&$75\\%$&II\\\\ & &0.2&$80\\%$&None& & &0.2&$10\\%$&None\\\\ 0.05 & 0.33 &0.0&$75\\%$&I&0.05 & 1 &0.0&$70\\%$&I\\\\ & &0.01&$75\\%$&I& & &0.01&$70\\%$&I\\\\ & &0.05&$65\\%$&II& & &0.05&$40\\%$&II\\\\ & &0.1&$40\\%$&II& & &0.1&$20\\%$&II\\\\ & &0.2&$5\\%$&None& & &0.2&$5\\%$&None\\\\ 0.05 & 2 &0.0&$60\\%$&I&0.05 & 3 &0.0&$55\\%$&I\\\\ & &0.01&$60\\%$&I& & &0.01&$55\\%$&I\\\\ & &0.05&$25\\%$&I& & &0.05&$15\\%$&I\\\\ & &0.1&$15\\%$&II& & &0.1&$10\\%$&None\\\\ & &0.2&$0\\%$&None& & &0.2&$0\\%$&None\\\\ \\hline \\multicolumn{10}{l}{$^a$In AU Myr$^{-1}$.}\\\\ \\multicolumn{10}{l}{$^b$In Jovian mass.}\\\\ \\multicolumn{10}{l}{$^c$Fraction of surviving planetesimals at the end of the simulation, i.e. $4$ Myr for $5$ AU Myr$^{-1}$ migration rate and $200$ Myr for $0.05$ AU Myr$^{-1}$.}\\\\ \\multicolumn{10}{l}{$^d$As in Fig. \\ref{resume}.}\\\\ \\end{tabular} \\end{table*} \\begin{table*} \\centering \\caption{\\label{complete_hot} Same as Table \\ref{normal} but for simulations with initially excited disks. The maximum initial eccentricity of the planetesimals is mentioned for all simulations. The migration rate is $0.5$ AU Myr$^{-1}$ for all simulations.} \\begin{tabular}{ccccc||ccccc} Max. ecc.$^a$ & Mass$^b$ & Ecc. & Surv. planetesimals$^c$ & Disk shape$^d$ &Max. ecc.$^a$ & Mass$^b$ & Ecc. & Surv. planetesimals$^c$ & Disk shape$^d$ \\\\ \\hline 0.1 & 0.035 &0.0&$100\\%$&None&0.1 & 0.05 &0.0&$100\\%$&II\\\\ & &0.01&$100\\%$&None& & &0.01&$100\\%$&None\\\\ & &0.05&$100\\%$&None& & &0.05&$100\\%$&None\\\\ & &0.1&$100\\%$&None& & &0.1&$100\\%$&None\\\\ & &0.2&$100\\%$&None& & &0.2&$55\\%$&None\\\\ 0.1 & 0.33 &0.0&$70\\%$&II&0.1 & 1 &0.0&$60\\%$&I\\\\ & &0.01&$70\\%$&II& & &0.01&$55\\%$&II\\\\ & &0.05&$85\\%$&II& & &0.05&$50\\%$&II\\\\ & &0.1&$70\\%$&None& & &0.1&$25\\%$&II\\\\ & &0.2&$20\\%$&None& & &0.2&$5\\%$&None\\\\ 0.1 & 2 &0.0&$50\\%$&I&0.1 & 3 &0.0&$40\\%$&I\\\\ & &0.01&$45\\%$&I& & &0.01&$40\\%$&I\\\\ & &0.05&$25\\%$&I& & &0.05&$15\\%$&I\\\\ & &0.1&$15\\%$&II& & &0.1&$10\\%$&I\\\\ & &0.2&$5\\%$&None& & &0.2&$0\\%$&None\\\\ 0.2 & 2 &0.0&$35\\%$&I&0.2 & 3 &0.0&$30\\%$&I\\\\ & &0.01&$35\\%$&I& & &0.01&$25\\%$&I\\\\ & &0.05&$20\\%$&I& & &0.05&$10\\%$&I\\\\ & &0.1&$10\\%$&I& & &0.1&$5\\%$&I\\\\ & &0.2&$5\\%$&None& & &0.2&$0\\%$&None\\\\ \\hline \\multicolumn{10}{l}{$^a$For planetesimals.}\\\\ \\multicolumn{10}{l}{$^b$In Jovian mass.}\\\\ \\multicolumn{10}{l}{$^c$Fraction of surviving planetesimals at the end of the simulation, i.e. $40$ Myrs.}\\\\ \\multicolumn{10}{l}{$^d$As in Fig. \\ref{resume}.}\\\\ \\end{tabular} \\end{table*}" }, "0801/0801.2372_arXiv.txt": { "abstract": "Two first order strongly hyperbolic formulations of scalar-tensor theories of gravity allowing nonminimal couplings (Jordan frame) are presented along the lines of the 3+1 decomposition of spacetime. One is based on the Bona-Mass\\'o formulation, while the other one employs a conformal decomposition similar to that of Baumgarte-Shapiro-Shibata-Nakamura. A modified Bona-Mass\\'o slicing condition adapted to the scalar-tensor theory is proposed for the analysis. This study confirms that the scalar-tensor theory has a well posed Cauchy problem even when formulated in the Jordan frame. ", "introduction": "\\label{sec:introduction} Scalar-tensor theories of gravity (STT) are alternative theories of gravitation where a scalar field is coupled nonminimally to the curvature associated with the {\\em physical metric}\\/ (this is the so-called {\\em Jordan frame}\\/ representation). The term ``physical metric'' refers to a situation where test particles follow the geodesics of that metric. The variation of the action of the STT with respect to the physical metric gives rise to field equations which contain an effective energy-momentum tensor (EMT) involving second order derivatives in time and space of the scalar field. Such EMT has the property that ``ordinary matter'', {\\em i.e.} matter associated with fields other that the scalar field, obeys the (weak) equivalence principle which mathematically translates into a conserved EMT for ordinary matter alone. Since {\\em a priori}\\/ it was not clear how such second order derivatives could be eliminated (in terms of lower order derivatives), or managed so as to obtain a quasilinear system of hyperbolic equations for which the Cauchy problem was well-posed (in the Hadamard sense), many people decided to abandon this approach in favor of the so-called {\\em Einstein frame}\\/ representation where the nonminimal coupling (NMC) is absorbed into the curvature by means of a conformal transformation of the metric. The new conformal metric is unphysical in the sense that (non-null) test particles do not follow the geodesics of that metric. However, the mathematical advantage is that the field equations for the nonphysical metric resemble the standard Einstein field equations with an unphysical effective EMT which involves at most first order derivatives of a suitable transformed scalar field (this EMT is unphysical because the ``ordinary matter'' part is not separately conserved). In the Einstein frame one can show that by using standard gauges ({\\em e.g.} harmonic gauge) the field equations acquire the form required in the application of theorems that establish the well-posedness of the Cauchy problem. In view of the apparent mathematical advantages and disadvantages of the Jordan and Einstein frames, several widespread misconceptions became common in the literature. One of these concerned the statement that the Cauchy problem is only well-posed in the Einstein frame~\\cite{Faraoni04}. In Ref.~\\cite{Salgado06}, however, one of us showed that this is not the case by following two different approaches. One was in the spirit of a second order analysis consisting on reducing the set of field equations, both for the metric components and the scalar field, to a manifestly quasilinear diagonal second order hyperbolic form (the ``reduced'' field equations). This was achieved by manipulating the field equations in a way that allowed one to express the d'Alambertian on the scalar field in terms of at most first order derivatives, and also by implementing a modified harmonic gauge which was adapted to the STT. Such a modified gauge allowed one to eliminate the remaining second order derivatives of the scalar field which had previously prevented the applicability of Leray's theorem (see {\\em e.g.} Ref.~\\cite{Wald84} for the theorem). The second approach followed in \\cite{Salgado06}, and which was directly related with the initial value problem, consisted in recasting the field equations of the STT in a 3+1 or Arnowitt-Deser-Misner (ADM)~\\cite{Arnowitt62} form written {\\em \\`a la} York~\\cite{York79} (hereafter referred to as the ADMY equations). To do so it was necessary to define suitable variables, followed by a manipulation of the equations in order to obtain well-defined constraints (independent of the choice of the lapse and shift), as well as first order in time evolution equations for the extrinsic curvature and the scalar-field variables. In Ref.~\\cite{Salgado06} the main goal was to focus in the 3+1 approach rather than the second order one because of the subsequent numerical applications we had in mind. In fact, almost all the modern codes used in numerical relativity are based on first order in time formulations, and therefore we wanted to adapt them for the analysis of phenomena in STT. Now, when taking the limit of pure general relativity (GR) ({\\em i.e.} no no-minimal coupling), the 3+1 equations presented in~\\cite{Salgado06} reduce to the standard ADMY equations with a minimally coupled scalar field plus ordinary matter sources. It is well-known that the ADMY equations are not strongly hyperbolic (see~\\cite{Alcubierre08a} for a detailed discussion), and therefore the corresponding equations for STT described in~\\cite{Salgado06} were expected not to be strongly hyperbolic either. As already mentioned in~\\cite{Salgado06}, such equations were only the first step towards a first order strongly hyperbolic system for which the well-posedness of the Cauchy problem could be established. The aim of this paper is then to fill that gap, and to obtain a first order strongly hyperbolic system of partial differential equations based on the 3+1 equations of~\\cite{Salgado06}. Actually, we will show here two such systems: one that is related to the Bona-Mass\\'o-Seidel-Stela (BMSS) formulation~\\cite{Bona94b} (which in turn is based on the Bona-Mass\\'o (BM) formulation~\\cite{Bona92}), and a second one based on the Baumgarte-Shapiro-Shibata-Nakamura (BSSN) conformal decomposition~\\cite{Baumgarte:1998te,Shibata95}. As was done in~\\cite{Salgado06}, a modified BM slicing condition will be used for the STT, while the shift vector will be taken as an {\\em a priori}\\/ given function of the coordinates. A last comment about the Jordan vs. Einstein frame is in order. In~\\cite{Salgado06} it was a matter of principle to show that the Cauchy problem could be well formulated in the Jordan frame. It then became clear that there was no fundamental reason to continue using the Einstein frame. After all, the Jordan frame is the one associated with the (physical) quantities that are to be confronted with the observations. Moreover, the constraint and evolution equations are not more involved than those of usual GR, so from a numerical point of view the use of the Jordan frame does not add much complexity to the analysis. On the other hand, the direct use of the Jordan frame allows a better physical interpretation of the results and avoids the potential problems that might arise in the Einstein frame in cases when the conformal transformations back to the physical metric are not well-defined. Finally, it has been argued that some of the energy densities associated with the Jordan frame effective EMT (JFEMT) are not positive definite and may therefore be unphysical, and also the corresponding ADM mass could then become negative. While it is certainly true that the JFEMT does not satisfy the energy conditions in general, it is very likely that any physical configuration consistent with observations will not carry (total) negative energy. This is because the deviations of any alternative theory from GR are presumably very small in order to reproduce many of the current observations (cf. Ref.~\\cite{Will93}). Actually, in any viable STT cosmology it turns out that the total energy density of the Universe is always positive, even when taking into account the negative contributions due to the NMC~\\cite{Salgado96,Salgado97a,Salgado97b,Quevedo99}. Again, this is because the scalar-field contributions (positive or negative) to the total energy density should be in agreement with several observations related to the past and present history of the Universe. For instance, the expansion rate of the Universe modified by the NMC contributions has to be consistent with the one required to produce the correct abundance of primordial nucleosynthesis~\\cite{Salgado96}. In addition, the density perturbations in STT should also match the Cosmic Microwave Background data. In other words, a STT cosmology which produces a total effective negative energy density of the Universe at any given epoch will surely not be consistent with observations. By the same token, in a consistent STT cosmology the negative contributions (if any) are to be naturally suppressed by the positive ones, providing a consistent Universe. In the case of astrophysical applications ({\\em e.g.} neutron star models) within a class of STT with a positive definite NMC function (which implies a positive definite effective gravitational ``constant''), the ADM mass turns to be always positive despite the negative contributions to the energy density due to the NMC~\\cite{Salgado98}. Moreover, the value of the NMC cannot be very high as otherwise the corresponding STT would put in jeopardy the agreement with the binary pulsar observations~\\cite{Damour96}. Of course, the nice thing about matter fields satisfying the energy conditions is the applicability of several theorems ({\\em e.g.} positive mass theorems and singularity theorems; see~\\cite{Wald84} and references therein). However, the fact that the JFEMT does not satisfy the energy conditions in general simply indicates that the theorems cannot say anything about the positivity of mass or the formation of singularities in this case. This is not a problem of physical but of mathematical character. But again, the nonpositive definite terms arising from the NMC will be surely bounded if the STT in hand is to be consistent with the current observations and experiments, and therefore the nice features of an EMT respecting the energy conditions will also very likely appear in any observationally consistent STT. Of course, all these arguments are only sustained by numerical experiments when constructing viable phenomenological models (in cosmology and compact objects), and therefore do not constitute a theorem. It would then be quite interesting to explore the possibility of proving positive energy and singularity theorems in the Jordan frame if one can bound the non-positive definite contribution associated with the JFEMT. This paper is organized as follows. In Sec.~\\ref{sec:ADM} we introduce the STT and the 3+1 equations described in~\\cite{Salgado06}. The notation of several variables will be slightly modified relative to Ref.~\\cite{Salgado06} in order to match with the one usually employed in numerical relativity. In Sec.~\\ref{sec:hyper} a fully first order strongly hyperbolic system of the field equations of STT is obtained along the lines of the BMSS and BSSN formulations of GR. We conclude with a discussion of future numerical applications of the hyperbolic systems presented here. Finally, in Appendix A we include the complete set of equations for both formulations, and in Appendix B we introduce a simple example to complement the ideas of Sec.~\\ref{sec:hyper}. ", "conclusions": "\\label{sec:discussion} In this paper we have constructed two novel first order strongly hyperbolic formulations of the STT in the Jordan frame along the lines of the BM and BSSN approaches. Such constructions show that both formulations have a well-posed Cauchy problem. This analysis fills the gap of a previous study on the Cauchy problem of STT~\\cite{Salgado06} and confirms that the Jordan frame is mathematically adequate for treating the initial value problem. One of the most interesting features of the formulations presented here is that a modified Bona-Mass\\'o slicing condition is required for the two new systems to be strongly hyperbolic while allowing several slicings ($f_{BM}>0$) which are natural generalizations of the slicings used in pure GR. In particular $\\Theta=1,f_{BM}$ in Eq.~(\\ref{STTBMlapse}) are two simple choices that lead to well behaved eigenfields. In the absence of a scalar field, the equations of Sec.IIIA,B reduce (for $\\varsigma=-1$, $\\zeta=4$ and $\\xi=2$) to the known BM and BSSN formulations of GR (when the NMC function is trivial, {\\em i.e.}, $F(\\phi)\\equiv 1$, the ``gravitational'' and the scalar-field sectors decouple completely up to principal part). What remains to be investigated is the usefulness and robustness of these formulations in actual numerical experiments, as well as the inclusion of a ``live shift''. Actually, we plan to analyze the dynamical transition to the phenomenon of spontaneous scalarization in boson stars arising in STT \\cite{Whinnett00} and the subsequent gravitational collapse to a black hole with gravitational wave emission of scalar type, using one or both of the hyperbolic formulations presented here. Both phenomena (spontaneous scalarization and scalar gravitational waves) are even present in spherical symmetry due to the NMC, therefore by assuming such a symmetry one can simplify the equations without eliminating the interesting physical features. \\bigskip An important consequence of the analysis presented here is that a slightly more general STT which includes a function $\\omega(\\phi)$ in the kinetic term of the scalar-field sector of the action~(\\ref{jordan}) ({\\em i.e.} the kinetic term has the form $\\omega(\\phi)(\\nabla \\phi)^2/2$) posses a well-posed Cauchy problem as well (except for some choices of $\\omega(\\phi)$; see Ref.~\\cite{Faraoni07}). The fact is that all the terms with $\\omega(\\phi)$ do not contribute to the principal part of the equations associated with the {\\em metric}\\/ sector ({\\em i.e.} the equivalent of Eq.~(\\ref{Einst})), while the scalar-field sector (the equivalent of Eq.~(\\ref{KG})) preserves the quasilinear diagonal hyperbolic form (see Ref.~\\cite{Faraoni07} for the detailed equations). Thus, up to principal part such STT are identical to the ones analyzed here. The relevance of this generalization is that such STT can be mapped to the so-called {\\em modified theories of gravity}\\/ which are given by a Lagrangian density $f(R)$ ($R$ being the Ricci scalar)~\\cite{Faraoni07}. Some specific choices of $f(R)$ lead to gravity theories which have been recently analyzed in several contexts. Notably, in the cosmological setting such theories have been proposed as an alternative to dark energy, since they can produce an accelerating expansion of the Universe without any exotic form of matter~\\cite{Carroll04,Capozziello03}. However, it must be emphasized that some of these theories might violate the solar system tests~\\cite{Erickcek06,Chiba07}, and some modifications are required to circumvent such drawbacks (see Ref.~\\cite{Nojiri07} for a review). The point we want to underline here is that, from the mathematical point of view, the viability of such theories relies heavily on the well-posedness of the Cauchy problem." }, "0801/0801.2975_arXiv.txt": { "abstract": "{}{ Several photometric redshift (photo-$z$) codes are discussed in the literature and some are publicly available to be used by the community. We analyse the relative performance of different codes in blind applications to ground-based data. In particular, we study how the choice of the code-template combination, the depth of the data, and the filter set influences the photo-$z$ accuracy.}{We performed a blind test of different photo-$z$ codes on imaging datasets with different depths and filter coverages and compared the results to large spectroscopic catalogues. { We analysed the photo-$z$ error behaviour to select cleaner subsamples with more secure photo-$z$ estimates.} We consider \\emph{Hyperz}, \\emph{BPZ}, and the code used in the CADIS, COMBO-17, and HIROCS surveys.} {{ The photo-$z$ error estimates of the three codes do not correlate tightly with the accuracy of the photo-$z$'s. While very large errors sometimes indicate a true catastrophic photo-$z$ failure, smaller errors are usually not meaningful.} For any given dataset, we find significant differences in redshift accuracy and outlier rates between the different codes { when compared to spectroscopic redshifts}. However, different codes excel in different regimes. { % The agreement between different sets of photo-$z$'s is better for the subsample with secure spectroscopic redshifts than for the whole catalogue. Outlier rates in the latter are typically larger by at least a factor of two.}} {{ Running today's photo-$z$ codes on well-calibrated ground-based data can lead to reasonable accuracy. The actual performance on a given dataset is largely dependent on the template choice and on realistic instrumental response curves. The photo-$z$ error estimation of today's codes from the probability density function is not reliable, and reported errors do not correlate tightly with accuracy. It would be desirable to improve this aspect for future applications so as to get a better handle on rejecting objects with grossly inaccurate photo-$z$'s. The secure spectroscopic subsamples commonly used for assessments of photo-$z$ accuracy may be biased toward objects for which the photo-$z$'s are easier to estimate than for a complete flux-limited sample, resulting in very optimistic estimates.}} ", "introduction": "\\label{sec:introduction} Photometric redshifts (hereafter, photo-$z$) have become a standard tool for the observing astronomer in the last years { \\citep{1986ApJ...303..154L, 1995AJ....110.2655C, 1999ASPC..191....3K, 1999A&A...343..399W, 2000ApJ...536..571B, 2000A&A...363..476B, 2001AJ....122.1151R, 2001AJ....122.1163B, 2001A&A...365..660W, 2003AJ....125..580C, 2003MNRAS.339.1195F, 2004PASP..116..345C, 2004MNRAS.353..654B, 2006A&A...457..841I, 2006MNRAS.372..565F}}. Not only are large multi-colour imaging surveys planned and executed with the goal of estimating the redshift of as many galaxies as possible from their broad-band photometry, but also many smaller projects benefit from this technique by providing redshifts that are much cheaper, in terms of telescope time, than spectroscopic ones and may go deeper. { Users of photo-$z$'s are often concerned with three main performance issues, which are the mean redshift error, the rate of catastrophic failures, and the validity of the probability density function (PDF) in a frequentist interpretation. The PDF may be correct in a Bayesian interpretation when including systematic uncertainties in the model fitting and correctly express a degree of uncertainty. However, given the non-statistical nature of systematic uncertainties a frequentist PDF that correctly describes the redshift distribution in the real experiment is necessarily different, unless such systematics can be excluded. The process depends on three ingredients: model, classifier, and data. A basic issue at the heart of problems with the PDF are the match between data and model, since best-fitting parameters and confidence intervals in $\\chi^2$-fitting are only reliable when the model is appropriate. The importance in choosing the type of data is the need to break degeneracies between ambiguous model interpretations. Finally, the classifiers are expected to produce similar results, while they could produce them at dramatically different speed. Artificial Neural Nets, hereafter ANNs, are especially fast once training has been accomplished \\citep{2003MNRAS.339.1195F}. There are many cases in the literature where the precision of photo-$z$'s has been improved after recalibrating the match between data and model \\citep[see e.g. ][]{2003AJ....125..580C,2004ApJS..150....1B,2006A&A...457..841I,2006AJ....132..926C,2007ApJS..172...99C}, although this process requires a large, representative training set of spectroscopic redshifts from the pool of data that is to be photo-$z$'ed. If ANNs are trained with sufficiently large training samples they can achieve the highest accuracies within the training range as a mismatch between data and model is ruled out from the start. The literature reports several different photo-$z$ estimators in use across the community, some of which use different template models and some of which allow implementation of user-defined template sets. Assuming a modular problem, where model (templates), classifier, and data can be interchanged, it is interesting to test how comparable the results of different combinations are. In this spirit, we have started the work presented in this paper, where we analyse photo-$z$ performance from real ground-based survey data, in dependence of magnitude, depth of data, filter coverage, redshift region, and choice of photo-$z$ code. We concentrate on the blind performance of photo-$z$'s which is the most important benchmark for any study that cannot rely on recalibration, e.g. in the absence of spectroscopic redshifts. We choose to focus on ground-based datasets because a lot of codes were tested on the Hubble-Deep-Field for which results can already be found in the literature \\citep[see e.g. ][]{1998AJ....115.1418H, 2000A&A...363..476B, 2000ApJ...536..571B}. Meanwhile, a much larger initiative has formed to investigate all (even subtle) differences in workings and outcomes among codes and models. This initiative called PHAT\\footnote{\\url{http://www.strw.leidenuniv.nl/~hendrik/PHAT}} (PHoto-$z$ Accuracy Testing) engages a world-wide community of photo-$z$ developers and users and will hopefully develop our understanding of photo-$z$'s to a reliably predictive level.} The paper is organised as follows. In Sect.~\\ref{sec:data} the imaging and spectroscopic datasets are presented. The photo-$z$ codes used for this study are described in Sect.~\\ref{sec:codes}. Sect.~\\ref{sec:strategy} presents our approach for describing photo-$z$ accuracy. The results are presented and discussed in Sect.~\\ref{sec:results}. { The different photo-$z$ estimates are compared to each other in Sect.~\\ref{sec:phz_vs_phz}.} A final summary and general conclusions are given in Sect.~\\ref{sec:conclusions}. Throughout this paper we use Vega magnitudes if not otherwise mentioned. ", "conclusions": "\\label{sec:conclusions} We have shown that photo-$z$'s estimated with today's tools can produce a reasonable accuracy. The performance of a particular photo-$z$ code, however, cannot easily be characterised by a mere two numbers such as scatter and global outlier rate. The benchmarks are rather sensitive functions of filter set, depth, redshift range and code settings. Moreover, there is at least a factor of two possible difference in performance between different codes which is again not stable for all setups but can vary considerably from one setup to another. There are, for example, redshift ranges where one code clearly beats another one in terms of accuracy only to loose at other redshifts. We give estimates of the performance for a number of codes in some practically relevant cases. The estimation of photo-$z$'s from different ground-based datasets is not straightforward and results should not be expected to be identical to simulated photo-$z$ estimates. Rather, photo-$z$ simulations often seem to circumvent critical steps in ground-based photo-$z$ estimation. { Most importantly}, the match between observed colours and some template sets commonly used may be suboptimal. In the preceding sections we have identified several aspects which are relevant to future optimisations of photo-$z$ codes. The photo-$z$ error estimation is one of the most unsatisfying aspects to date with error values often only very weakly correlated with real uncertainties. This is likely due to the insufficient inclusion of systematics since very low S/N objects, for which the errors should be dominated by photon shot-noise, show a tighter correlation. Chip-to-chip sensitivity variations, especially in the UV, could either be taken into account more accurately within the photo-$z$ codes or could be tackled by improved instrument design, survey strategy, and data reduction. The optimisation of template sets can be expected to be successfully done with ever larger spectroscopic catalogues becoming available. { In general, biases can be removed by a recalibration which requires an extensive spectroscopic training set. Another proven successful route to better photo-$z$'s is improving the spectral resolution of the data, instead of their depth, as demonstrated by the COMBO-17 survey. This approach is also taken by the new ALHAMBRA survey \\citep[][ Ben\\'itez et al., 2007, A\\&A, submitted]{2005astro.ph..4545M} and COSMOS-21. A general problem for all studies comparing photo-$z$'s to spectroscopic redshifts is our finding that secure spectroscopic samples can be biased. While surveys like VVDS are $>90\\%$ complete in obtaining spectra for galaxy samples the redshifts that are claimed to be $>90\\%$ secure only amount to $\\sim50\\%$. This subsample obviously consists of galaxies for which the photo-$z$ estimation works better than for the whole sample. In the future, it is desirable to put effort into spectroscopic surveys with secure redshift measurements for virtually every galaxy down to the same flux limit that is used for the analysis of photo-$z$ samples. Several questions that are raised in this work will be tackled by the PHAT initiative mentioned above. PHAT aims to understand the issues presented here in a systematical and quantitative way in order to give guidance for better photo-$z$'s in the future.}" }, "0801/0801.1147_arXiv.txt": { "abstract": "A two-dimensional electrodynamic model is used to study particle acceleration and non-thermal emission mechanisms in the pulsar magnetospheres. We solve distribution of the accelerating electric field with the emission process and the pair-creation process in meridional plane, which includes the rotational axis and the magnetic axis. By solving the evolutions of the Lorentz factor, and of the pitch angle, we calculate spectrum in optical through $\\gamma$-ray bands with the curvature radiation, synchrotron radiation, and inverse~-Compton process not only for outgoing particles, but also for ingoing particles, which were ignored in previous studies. We apply the theory to the Vela pulsar. We find that the curvature radiation from the outgoing particles is the major emission process above 10~MeV bands. In soft $\\gamma$-ray to hard X-ray bands, the synchrotron radiation from the ingoing primary particles in the gap dominates in the spectrum. Below hard X-ray bands, the synchrotron emissions from both outgoing and ingoing particles contribute to the calculated spectrum. The calculated spectrum is consistent with the observed phase-averaged spectrum of the Vela pulsar. Taking into account the predicted dependency of the emission process and the emitting particles on the energy bands, we compute the expected pulse profile in X-ray and $\\gamma$-ray bands with a three-dimensional geometrical model. We show that the observed five-peak pulse profile in the X-ray bands of the Vela pulsar is reproduced by the inward and outward emissions, and the observed double-peak pulse profile in $\\gamma$-ray bands is explained by the outward emissions. We also apply the theory to PSR B1706-44 and PSR B1951+32, for which X-ray emission properties have not been constrained observationally very well, to predict the spectral features with the present outer gap model. ", "introduction": "The Compton Gamma-Ray Observatory (CGRO) had measured puled $\\gamma$-ray emissions from younger pulsars (Thompson 2004). The multi-wavelength observations of the $\\gamma$-ray pulsars have shown that the spectra of the non-thermal emissions extend in $\\gamma$-ray through optical bands. The observations have also revealed the pulse profiles in optical through $\\gamma$-ray bands, and the polarization properties of the pulsed optical emissions from the $\\gamma$-ray pulsars (Kanbach et al. 2005; Mignani et al 2007). In the future, furthermore, the polarization of the X-rays and $\\gamma$-rays from the pulsars will probably be able to be measured by ongoing projects (Kamae et al. 2007; Chang et al. 2007). The multi-wave length observations on the spectrum, the pulse profile and the polarization allow us to perform a comprehensive theoretical discussion for the mechanisms of the particle acceleration and of the non-thermal emission in the pulsar magnetospheres. The particle acceleration and the non-thermal emission processes in the pulsar magnetosphere have been mainly argued with the polar cap accelerator model (Sturrock 1971; Ruderman \\& Sutherland 1975) and the outer gap accelerator model (Cheng, Ho \\& Ruderman 1986a, 1986b). Also, the slot gap model (Muslimov and Harding 2004), which is an extension model of the polar cap model, was proposed. All models predict an acceleration of the particles by an electric field parallel to the magnetic field. In the pulsar magnetosphere, the accelerating electric field arises in a charge depletion region from the so called Goldreich-Julian charge density (Goldreich \\& Julian 1969), and the strong acceleration region in the magnetosphere depends on th e model. The polar cap and the slot gap models predict a strong acceleration within several stellar radii on the polar cap, and the outer gap model predicts a strong acceleration beyond the null charge surface, on which the Goldreich-Julian charge density becomes zero. As CGRO had observed, most of the $\\gamma$-ray pulsars have a double peak structure in the pulse profile in $\\gamma$-ray bands. The outer gap model has been successful in explaining the observed double peak structure (Romani \\& Yadigaroglu 1995). With the outer gap model, the double peak structure is naturally produced as an effect of the aberration and the time delay of the emitted photons by the outgoing particles that move toward the light cylinder, which is defined by the positions, on which the rotating speed with the star is equal to speed of the light. The axial distance of the light cylinder is $R_{lc}=c/\\Omega$, where $c$ is the speed of the light, and $\\Omega$ is the angular velocity of the neutron star. Because the Crab pulsar has the pulse profiles with the double peak structure in whole energy bands, the outer gap model can explain the pulse profiles of the Crab pulsar in optical through $\\gamma$-ray bands (Takata \\& Chang 2007). Furthermore, the outer gap model can explain the observed spectra, and the polarization characteristics in optical bands for the Crab pulsar (Cheng et al. 2000; Hirotani 2007; Jia, et al 2007; Takata et al, 2007; Takata \\& Chang 2007; Tang et al. 2007). Therefore, the outer gap accelerator model has successfully explained the results of the multi-wave length observations for the Crab pulsar. For the Vela pulsar, multi-peak structure in pulse profiles in X-ray, UV and optical bands have been revealed (Harding et al. 2002; Romani et al. 2005), although the pulse profile in $\\gamma$-ray bands has two peaks in a single period (Fierro et al. 1998). For example, Harding et al. (2002) analysed RXTE data and revealed the pulse profile, which has at least five peaks in a single period. Because the two peaks in the five peaks are in phase with the two peaks in the $\\gamma$-ray bands, the two peaks will be explained by the traditional outer gap model with the outward emissions from the outer gap. However, because other three peaks are not expected by the traditional outer gap model, the origin of the three peaks has not been understood, so far. The inward emissions emitted by particles accelerated toward stellar surface will give one possibility to explain the unexpected three peaks of pulse profile in the X-ray bands of the Vela pulsar. Because the primary particles are produced near the inner boundary of the gap, the ingoing particles feel a small part of whole potential drop in the gap before escaping the gap from the inner boundary, although the outward moving particles can feel whole potential drop before escaping the gap from the outer boundary. In previous studies, therefore, the contribution of the inward emissions on the spectral calculation has been ignored by assuming that its flux is much smaller than that of the outward emissions. However, it will be true only in the $\\gamma$-ray bands with the curvature radiation. Because main emission mechanism in the X-ray bands will be the synchrotron radiation, the inward synchrotron emissions with a stronger magnetic field will be efficient enough to contribute on the observed emissions below $\\gamma$-ray bands. Although the inward emissions have been consistently dealt in the electrodynamic studies (Takata et al. 2006; Hirotani, 2007), only the outward emissions were taken into account for computing the spectrum. In this paper, we calculate the spectrum in optical through $\\gamma$-ray bands by taking into account both inward and outward emissions. Specifically, we solve the electrodynamics in the outer gap accelerator in the two-dimensional plane, which includes the rotational axis and the magnetic axis. We calculate the spectrum of the curvature radiation, synchrotron radiation, and inverse~-Compton process for the primary and the secondary particles. We also discuss the expected pulse profiles from optical through $\\gamma$-ray bands with a three-dimensional model for comparison with the observations. Furthermore, we apply the theory to PSRs B1706-44 and B1951+32 to calculate the spectrum, because the properties of optical and X-ray emissions from the two pulsars have not been understood well. In section~2, we describe our two-dimensional electrodynamic model following Takata et al. (2004, 2006). In section~3, we apply the theory to the Vela pulsar, and we discuss the spectrum and the pulse profile in optical through $\\gamma$-ray bands. We also show the expected spectra of PSRs B1706-44 and B1951+32. ", "conclusions": "\\subsection{Phase resolved-spectrum} With the RXTE observations, Harding et al. (2002) proposed the spectral index $s$, which is defined by $I_{\\nu}\\propto \\nu^{-s}$, of $s\\sim 1$ for the phase-resolved spectrum of the peak (Pk~2-soft in figure~1 of Harding et al, 2002), which appears between the two peaks, whose phases are aligned with the peak positions in $\\gamma$-ray bands. As we discussed in section~\\ref{pulsep}, the preset model explains emission mechanism of RXTE Pk~2-soft with the synchrotron emissions of the ingoing particles, which create a peak (Figure~\\ref{pulse}, IP1) in the pulse profile. Although the spectrum in Figure~\\ref{spe1} is calculated using two-dimensional model, we may be able to read the spectral index of the phase-resolved spectra of the peak IP1 from Figure~\\ref{spe1} ( from the slope of the thin dashed-dotted line in the right panel), which predicts the spectral index $s\\sim 0$ for the IP1 emissions. For the reason of this discrepancy between the spectral indexes of RXTE and the present model, we argue that the three-dimensional structure effects the phase-resolved spectrum, and/or that RXTE observations might not determine the spectral behavior in the X-ray bands very well due to a bright synchrotron X-ray nebula. The present model predicts the inward synchrotron emissions for the ingoing primary particles (Figure~\\ref{spe1}, the thin dashed line in the left panel) contribute to the spectrum in hard X-ray and soft $\\gamma$-ray bands. Therefore it is important to measure the phase-resolved spectra and the pulse profiles of the Vela pulsar in the hard X-ray and soft $\\gamma$-ray bands to see how the phase-resolved spectra and the pulse profiles evolve. \\subsection{Correlation between radio emission and the outer gap emission} Recently, the correlation between the arrival time of the radio emissions and the shape of the non-thermal X-ray pulse profile of the Vela pulsar was discovered (Lommen et al. 2007). Although the radio emission mechanism has not been understood well, the polarization measurement indicates a correlation between the magnetic field geometry of the polar cap region and the swing of the position angle of the polarization of the radio emissions. Therefore, the polar cap accelerator model as the origin of the radio emissions from the pulsar has been widely accepted. A magnetospheric model having both the polar cap accelerator and the outer gap accelerator has been proposed upon request to balance between the energy loss rate and the angular momentum loss rate, which is equal to the energy loss rate divided by the angular velocity of the star, of the global magnetosphere (Shibata 1991; 1995). Although the polar cap accelerator and the outer gap accelerator can not exist on the same magnetic field lines because of the different current directions in the polar cap and outer gap accelerators, the polar cap accelerator affects the outer gap accelerator to obtain the torque balance. According to this model, therefore, even though the non-thermal X-rays originate from the outer gap accelerator, some correlations between X-ray emissions and the radio emission are expected. Furthermore, the present model also predicts that the outer gap accelerator affects the polar cap accelerator with the inward emissions, which pass near the stellar surface. The inward emissions will affect the circumstance around polar cap region through the pair-creation process and/or the scattering process. \\subsection{Inward emissions from the Crab pulsar} \\label{crab} Unlike the Vela pulsar, the Crab pulsar has only two peaks in the pule profile in optical through $\\gamma$-ray bands. This indicates that the observed emissions of the Crab pulsar are dominated by the outward emissions in whole energy bands. For the Crab pulsar, the photons above 1~GeV emitted via the curvature radiation in the outer gap are converted into the pairs outside the gap by the pair-creation process by the magnetospheric X-ray photons emitted by the secondary pairs. Outer gap model predicts the synchrotron and the inverse~-Compton processes of the secondary pairs explain the observed emissions in optical through $\\gamma$-ray bands. As we discussed in section~\\ref{spectrum} (Figure~\\ref{spe1},right), the outward synchrotron radiation by the outgoing secondary pairs created by the magnetospheric X-rays is much brighter than the inward synchrotron radiation, because the outgoing $\\gamma$-rays emitted by the curvature radiation is much more than the ingoing $\\gamma$-rays. For the Crab pulsar, therefore, we can expect that the outgoing secondary pairs produce the observed spectrum with the synchrotron and the inverse~-Compton process. Because the flux of the outward emissions of the secondary particles is much (say one or two order) larger than that of the inward emissions, the pulse profiles have strong two peaks in a single period in optical through $\\gamma$-ray bands for the Crab pulsar. Although we expect contribution of the inward emissions at off-pulse phase, it must be difficult to separate the tiny flux of the inward emissions from off-pulse outward emissions inside null charge surface and/or from the strong background emissions from the synchrotron nebula." }, "0801/0801.1692_arXiv.txt": { "abstract": "We have used archival \\chandra\\ and \\xmm\\ observations of quasars hosting intrinsic narrow UV absorption lines (intrinsic NALs) to carry out an exploratory survey of their X-ray properties. Our sample consists of three intrinsic-NAL quasars and one ``mini-BAL'' quasar, plus four quasars without intrinsic absorption lines for comparison. These were drawn in a systematic manner from an optical/UV-selected sample. The X-ray properties of intrinsic-NAL quasars are indistinguishable from those of ``normal'' quasars. We do not find any excess absorption in quasars with intrinsic NALs, with upper limits of $\\nh \\ls{\\rm a\\; few}\\times 10^{22}~\\cmm$. We compare the X-ray and UV properties of our sample quasars by plotting the equivalent width and blueshift velocity of the intrinsic NALs and the X-ray spectral index against the ``optical-to-X-ray'' slope, \\aox. When BAL quasars and other AGNs with intrinsic NALs are included, the plots suggest that intrinsic-NAL quasars form an extension of the BAL sequences and tend to bridge the gap between BAL and ``normal'' quasars. Observations of larger samples of intrinsic-NAL quasars are needed to verify these conclusions. We also test two competing scenarios for the location of the NAL gas in an accretion-disk wind. Our results strongly support a location of the NAL gas at high latitudes above the disk, closer to the disk axis than the dense BAL wind. We detect excess X-ray absorption only in Q0014+8118, which does not host intrinsic NALs. The absorbing medium very likely corresponds to an intervening system at $z=1.1$, which also produces strong absorption lines in the rest-frame UV spectrum of this quasar. In the appendix we discuss the connection between UV and X-ray attenuation and its effect on \\aox. ", "introduction": "\\label{sec:intro} The {\\it intrinsic} absorption lines found in the rest-frame UV spectra of quasars and active galactic nuclei (AGNs) are classified based on their line widths into broad absorption lines (BALs; FWHM$\\,>2000$~\\kms, found in $\\sim 10\\%$ of all quasars), narrow absorption lines (NALs; FWHM$\\,< 500$~\\kms, found in $\\sim 50\\%$ of quasars at $z\\sim 2$--4; these can be clearly separated from intervening lines via high-resolution spectroscopy), and mini-BALs (intermediate FWHM between BALs and NALs). They are typically blueshifted relative to the quasar and are thought to trace outflows from the quasar central engine. These outflows could be hydromagnetic accretion-disk winds (e.g., Blandford \\& Payne 1982; Emmering, Blandford \\& Shlosman 1992; K\\\"onigl \\& Kartje 1994; Everett 2005), or accretion-disk winds driven by radiation pressure (e.g., Murray et al. 1995; Arav, Li, \\& Begelman 1994, Proga, Stone, \\& Kallman 2000). Alternatively they could be outflows driven by thermal pressure and launched either from the accretion disk itself (e.g., Begelman, McKee, \\& Shields 1983) or from the obscuring torus invoked by AGN unification schemes and driven by thermal pressure (e.g., Balsara \\& Krolik 1993; Krolik \\& Kriss 1995, 2001; Chelouche \\& Netzer 2005). However, thermal accretion disk winds are too hot to produce absorption lines in the UV, while thermal winds from the obscuring torus are too slow to account for the observe velocities of NALs and BALs. Accretion disk winds, whatever their origin, are an integral part of the accretion flow, and may be the source of the broad emission lines that are characteristic of the optical and UV spectra of quasars and AGNs (e.g., Shields 1977; Chiang \\& Murray 1996; Murray \\& Chiang 1997). Moreover, the outflows may be important for cosmology since they deliver energy and momentum to the interstellar and intergalactic media (hereafter, ISM and IGM, respectively) and can affect galaxy evolution (e.g., Granato et al. 2004; Scannapieco \\& Oh 2004; Springel, Di Matteo \\& Hernquist 2005; Chartas et al. 2007a). Most of our knowledge about intrinsic absorbers is derived from optical/UV observations. A combination of observational results and models suggest that intrinsic absorption lines of different widths represent either different lines of sight through the outflowing wind to the quasar continuum source (Ganguly et al. 2001; Elvis 2000), or different stages in the evolution of the absorbing gas parcels (e.g., Hamann \\& Sabra 2004; Misawa et al. 2005). X-ray observations of BAL quasars have revealed large columns of nearly-neutral absorbing gas ($\\sim 10^{23-24}$~\\cmm; see, for example, Green \\& Mathur 1996; Gallagher et al. 2002a). However, little is known about the X-ray properties of quasars with intrinsic NALs and mini-BALs. X-ray spectroscopy of the last two types of quasars can be extremely useful in many respects: \\begin{itemize} \\item In general terms, we can compare the X-ray properties of quasars that host intrinsic NALs with those that do not. This comparison allows us to assess whether there is a significant difference between the central engines of the two types of object. If no difference is found, and since intrinsic NALs are ubiquitous in quasar spectra (see Misawa et al 2007b and references therein), we will be led to a picture where intrinsic NALs, hence outflows, are a universal property of quasars. \\item X-ray spectra can probe highly-ionized media that are not directly probed by optical/UV absorption lines. Moreover, they can yield more direct measurements of the total hydrogen column density of absorbers in a relatively low-ionization state (this cannot be easily done through UV absorption lines, unless the ionization state of the gas is very well constrained). Thus, we can use the X-ray spectra to investigate whether any hot gas (which may represent the bulk of the mass) co-exists (mixed or layered) with the UV absorbers. \\item The column densities determined from the X-ray spectra constitute a direct test of scenarios for the location of the NAL gas in the larger outflow. In particular, Elvis (2000) suggests that the NAL gas is located at very low latitudes above the accretion disk, implying very large column densities (comparable to those found in BAL quasars, i.e., $\\gs 10^{22}$~\\cmm). On the other hand, Ganguly et al. (2001) place the NAL gas at high latitudes above the accretion disk, implicitly suggesting lower column densities. \\item Several examples of intrinsic NALs at high ejection velocities ($\\sim 0.2c$) are now known (see Misawa et al. 2007b and references therein). It is extremely interesting, therefore, to search quasars with intrinsic NALs for high-velocity X-ray absorption lines (see, for example, Chartas et al. 2002, 2003). Such observations can yield results relevant to cosmology since they can lead to estimates of the mass outflow rate and the kinetic power of the outflow (see, for example, Chartas et al. 2007b). \\end{itemize} With the above considerations in mind, we have used archival \\chandra\\ and \\xmm\\ data to carry out an exploratory survey of the X-ray properties of a small, but carefully-selected sample of quasars at $z\\approx 2.5$--3.8. This sample consists of three quasars hosting intrinsic NALs, one quasar hosting a mini-BAL (HS1603+3820), and four quasars without intrinsic absorption lines for comparison (these were selected in a systematic way, as described in the next section). We compare the properties of the UV NALs (e.g., equivalent width, outflow velocity) with the parameters describing the X-ray spectrum (e.g., photon index, {\\it intrinsic} column density, and optical-to-X-ray slope, \\aox). Thus, we place NAL quasars in the context of BAL quasars (see Gallagher et al. 2002b; Brandt, Laor, \\& Wills 2000) and investigate whether all types of objects follow the same trends in their properties. We also use the column densities determined from the X-ray spectra to carry out one of the tests outlined, above, namely, we attempt to distinguish between the two different suggestions for the location of the NAL gas in the larger outflow. In \\S\\ref{sec:sample}, we describe how the sample was selected and summarize the properties of the constituent quasars. In \\S\\ref{sec:obs} and \\S\\ref{sec:analysis}, we present the analysis of the data (observations, data screening, and model fits to the X-ray spectra). We compare the rest-frame UV and X-ray properties in \\S\\ref{sec:xuv}. In \\S\\ref{sec:discussion} we summarize our findings, discuss our results, and consider prospects for future work. We adopt a cosmological model with $H_{0}$ = 75~\\kms~Mpc$^{-1}$, $\\Omega_{m}$=0.3, and $\\Omega_{\\Lambda}$ = 0.7. Throughout this paper we give error bars corresponding to the 90\\% confidence level, unless noted otherwise. A significant part of our discussion makes use of the optical-to-X-ray slope of the spectral energy distribution, \\aox, which is affected by extinction. Thus we examine the effects of extinction on \\aox\\ in detail in the appendix. ", "conclusions": "\\label{sec:discussion} The results presented above show that there are no differences in X-ray properties between quasars with and without intrinsic NALs in our sample. The spectra of all the quasars in our sample can be described by a very simple model, namely a power law modified by absorption by nearly-neutral matter at the redshift of the source. In all but one quasar, we are only able to obtain upper limits to the column density of the intrinsic X-ray absorber; these limits are of order a few $\\times 10^{22}$~\\cmm. The possibility that the intrinsic X-ray absorbers are ionized can be neither confirmed nor ruled out based on the data we have analyzed here. The only quasar in which we have found a measurable column density is Q0014+8118 ($\\nh\\approx 1\\times 10^{22}$~\\cmm), which does {\\it not} host any intrinsic NALs. We discuss this quasar further below. Our comparison with larger samples of quasars of comparable luminosity and redshift shows that quasars with intrinsic NALs do not differ from the general population in either their X-ray spectra or their spectral energy distributions as quantified by \\aox. The rest-frame equivalent widths of intrinsic NALs follow the same trend with \\aox\\ as nearby quasars and Seyfert galaxies, suggesting a relation between the medium responsible for the intrinsic NALs and the medium responsible for the X-ray continuum absorption. However, X-ray observations of a larger sample of quasars with intrinsic NALs is sorely needed to strengthen our conclusions. A conclusion that follows immediately from the above results is that outflows that manifest themselves in the form of intrinsic NALs could be present in all quasars. We showed in our earlier work that intrinsic NALs are very common (occurring in 50\\% of all quasars at $z\\sim 2$--4; Misawa et al. 2007b) and the results presented here indicate that the typical spectral characteristics of quasars in general do not preclude their hosting intrinsic NALs. \\begin{figure} \\centerline{\\includegraphics[width=3.4in]{f6.eps}} \\caption{Comparison of the properties of quasars in our sample with the properties of BAL quasars. Quasars from our sample with intrinsic NALs are represented by stars, while those without are represented by open stars. For the two quasars that were observed more than once (Q1422$+$2309 and Q1700$+$6416), we plot a single, representative value of \\aox\\ (see discussion in \\S\\ref{sec:xuv} of the text). Filled circles represent BAL quasars from the sample of Gallagher et al. (2006). {\\it Top Panel:} Variation of the total rest-frame equivalent width (summed over all intrinsic NALs in the same quasar) with \\daox. The rest-frame equivalent width assigned to quasars without intrinsic NALs corresponds to the typical observed-frame detection limit of 0.056~\\AA\\ (see Misawa et al 2007b). {\\it Middle Panel:} Variation of X-ray photon index with \\daox. Open triangles in this panel represent the extremely red quasars from Hall et al. (2006). {\\it Bottom Panel:} Variation of maximum blueshift velocity of intrinsic NALs with \\daox. One intrinsic NAL quasar from our sample (Q0130$-$4021; $\\vmax\\approx 65,000~\\kms$) is off scale and marked with upward arrow.\\label{fig:correlations}} \\end{figure} In the remainder of this section, we compare the properties of quasars with intrinsic NALs from our sample with those of BAL quasars studied by other authors, we use our results to constrain scenarios for the geometry and location of the intrinsic NAL gas, and we discuss the properties of the absorber in Q0014+8118. We conclude by noting open questions and possible directions for future work. \\subsection{Comparison with BAL quasars}\\label{sec:comp} We compare the properties of the NAL quasars in our sample with those of BAL quasars in the sample of Gallagher et al. (2006) and very red quasars from Hall et al. (2006); their redshifts are between 1.4 and 2.8 and their luminosities are comparable to those of our quasars. In Figure~\\ref{fig:daox}, we show the distribution of {\\it measured} values of \\daox\\ among BAL quasars on the same scale as the distribution of \\daox\\ among quasars in our sample. There is an obvious difference between the two distributions; a Kolmogorov-Smirnov test yields a chance probability of 0.003 that the intrinsic NAL quasars in our sample and the BAL and very red quasars from the bottom panel of Figure~\\ref{fig:daox} were drawn by chance from the same parent population. This suggests that the X-ray source in intrinsic NAL quasars is not as heavily obscured as in BAL quasars (see the discussion of the \\daox\\ distribution in BAL quasars by Gallagher et al. 2006). In Figure~\\ref{fig:correlations}, we plot properties of UV BALs as well as the X-ray photon index for the Gallagher et al. (2006) sample against \\daox. In the same plots we include the quasars from our own sample for comparison. The top two panels of Figure~\\ref{fig:correlations} suggest that quasars with intrinsic NALs fit in a progression between BALs and ``normal'' quasars. As the rest-frame equivalent width of the \\ion{C}{4} absorption line decreases, \\daox\\ increases, in a manner that is qualitatively consistent with a decrease in the column density of the absorber. This picture is supported by the photon index vs \\daox\\ plot, where quasars with intrinsic NALs occupy the low-column-density end of the BAL distribution. It is reassuring that the mini-BAL quasar in our sample, HS1603+3820, falls closer to the BAL locus in both of the top two panels of Figure~\\ref{fig:correlations} (as well as in the top panel of Fig.\\ref{fig:compaox}) than the intrinsic NAL quasars. In contrast, in the plot of maximum blueshifted velocity vs \\daox\\ shown in the bottom panel of Figure~\\ref{fig:correlations}, quasars with intrinsic NALs do not fit into the BAL sequence, even though this sequence appears to connect smoothly to the ``normal'' quasars. This behavior can be reconciled with the behavior seen in the top two panels of the same figure in the context of an equatorial wind scenario. In the top two panels of Figure~\\ref{fig:correlations} the BAL quasar tracks are defined primarily by the column density of the absorbing medium since $\\log\\wrest$, $\\Gamma$, and \\daox\\ are all measures of the column density (the values of $\\Gamma$ of BAL quasars were inferred from the hardness ratio, thus they are directly affected by absorption and so are the 2~keV flux densities that are based on the same data). On the other hand, the orientation of the outflow relative to the line of sight and the acceleration mechanism play an important role in defining the BAL quasar track in the bottom panel of this figure, since they influence the value of \\vmax\\ (see for example, the discussion in Gallagher et al 2006). Therefore, the deviation of intrinsic NAL quasars from the BAL track suggests that the NAL gas is not part of the dense BAL flow thought to be located near the base of the wind, although it may still be associated with it. In other words, our line of sight through the NAL gas does not generally pass through the dense BAL gas. This conclusion also serves as a test of scenarios for the location of the NAL gas within the greater outflow, as we explain further in \\S\\ref{sec:geom}, below. \\subsection{Geometry and Location of Absorbing Medium}\\label{sec:geom} Using our observational results, we can test and constrain scenarios for the location of the intrinsic NAL gas in an (equatorial) accretion-disk wind whose dense, low-latitude parts are likely to give rise to BALs. The two competing scenarios are depicted in Figures~1--3 of Elvis (2000) and in Figure~13 of Ganguly et al. (2001). Elvis (2000) proposed that quasars with intrinsic NALs are viewed at a large inclination angle relative to the disk/wind axis (larger than in the case of BALs). As a result the column density through the NAL line of sight is $\\nh\\gs 10^{22}$~\\cmm\\ and the apparent blueshift of the intrinsic NALs should be considerably smaller than that of BALs, ($\\voff\\sim10^3$~\\kms). On the other hand, in the Ganguly et al. (2001) picture the NAL gas is located {\\it above} the BAL gas, closer to the axis of the disk/wind. The column density is lower because the NAL gas is distributed in small parcels and the blueshifts can be as high as those found in BALs (or even higher). Our results favor the Ganguly et al. (2001) scenario for a number of reasons. First, we do not detect large absorbing column densities; we find that $\\nh\\ls {\\rm a\\; few}\\times 10^{22}$~\\cmm\\ in all of our quasars. Second, the observed blueshifts of intrinsic NALs are comparable to or higher than those of BALs, suggesting that the absorbing gas is not associated with the base of the wind. Rather, this gas should have traveled an appreciable distance from its launch point and attained a speed close to its terminal speed (see, for example, Murray et al. 1995 and Hamann 1998). Third, the comparison of intrinsic NALs and BALs presented above (\\S\\ref{sec:comp} and bottom panel of Fig.~\\ref{fig:correlations}) shows that the intrinsic NALs do not connect smoothly to the BAL sequence in the \\vmax--\\daox\\ diagram, suggesting that the NAL gas is not a part of the dense BAL flow. In contrast, the Ganguly et al. scenario passes these tests and also places the NAL gas far enough away from the central continuum source so that (a) it is not highly ionized, and (b) it can be accelerated to a high speed. \\subsection{The Case of Q0014+8118}\\label{sec:case} Q0014+8118 is the only quasar in which we have found excess absorption but it is also a quasar not known to have intrinsic UV NALs. We explore here whether these two observational results can be reconciled. We have re-examined the rest-frame UV spectrum of this quasar presented in Misawa et al. (2007b) and found no strong absorption lines near the redshift of this quasar, which could be attributed to a neutral absorber with the column density comparable to that measured in the X-ray spectrum. We did, however, find an intervening absorption line system at $z=1.1$ with strong \\ion{Fe}{2}$\\;\\lambda$2600, \\ion{Mg}{2}$\\;\\lambda\\lambda$2796,2803 and \\ion{Mg}{1}$\\;\\lambda$2853 absorption lines with the following rest-frame equivalent widths: $W_{rest}(\\lambda 2600)=2.06\\;$\\AA, $W_{rest}(\\lambda 2796)=3.23\\;$\\AA, and $W_{rest}(\\lambda 2853)=0.67\\;$\\AA. This raises the possibility that the excess X-ray absorption is associated with the lower-redshift intervening system. We consider each of these two scenarios in turn. To assess whether the X-ray absorber could be intrinsic to the quasar, we have carried out a number of photoionization simulations using the code Cloudy (Ferland et al. 1998) adopting the Mathews \\& Ferland (1987) model for the quasar spectral energy distribution (but modified to match the X-ray photon index of Q0014+8118), a Solar metallicity and abundance pattern, the total Hydrogen column density determined from the X-ray spectrum, and two different values of the density ($10^4$ and $10^8$~\\cmmm; the results turn out not be sensitive to the density). We computed four models with ionization parameters corresponding to $\\log\\xi_{\\rm XSTAR}=1.04$, 1.84, 2.24, and 2.54, which span the range of values inferred for the X-ray absorber (see \\S\\ref{sec:notes}). In all cases, the equivalent widths of the \\ion{C}{4}, \\ion{N}{5}, and \\ion{O}{6} absorption lines are large enough that these lines should have been detected in the high-resolution spectrum presented in Misawa et al. (2007b). Missing these lines would require a rather unlikely set of circumstances, namely: the ionization parameter should be at the high end of the allowed range so that the lines are weak and the outflow velocity should be such that the \\ion{C}{4} line is out of the range of the observed UV spectrum, the \\ion{N}{5} line should fall in the gap between echelle orders, and the \\ion{O}{6} line should be hidden in the Ly$\\alpha$ forest. Therefore, we disfavor the intrinsic interpretation of the absorber. To assess the intervening absorber hypothesis we make use of the statistical observational results of Rao, Turnshek, \\& Nestor (2006). According to these authors, an intervening absorber with the observed \\ion{Fe}{2}, \\ion{Mg}{2}, and \\ion{Mg}{1} equivalent widths has a 65\\% probability of being a damped Ly$\\alpha$ system (DLA), with an average hydrogen column density of $4\\times 10^{20}$~\\cmm\\ and a column density of $1\\times 10^{21}$~\\cmm\\ being reasonably likely. Moreover, the observed relative strengths of the \\ion{Fe}{2} and \\ion{Mg}{2} lines suggest that the hydrogen column density can plausibly be as high as $5\\times 10^{21}$~\\cmm, while the observed \\ion{Mg}{1} strength suggests a 50\\% probability of $N_{\\rm H} > 1\\times 10^{21}$~\\cmm. Thus we favor the intervening absorber hypothesis because the known intervening system at $z=1.1$ has a high probability of producing just the X-ray absorption we observe. In this context we may also understand the low attenuation of the rest-frame UV light of this quasar. By comparing the colors of large samples quasars with and without DLAs, Ellison, Hall, \\& Lira (2005) find that the dust content of DLAs is very small and place a limit of $E(B-V)<0.04$ on the color excess that they produce. \\subsection{Open Questions and Future Prospects}\\label{sec:prospects} There are a number of outstanding questions that our exploratory survey was not able to address. \\begin{itemize} \\item We have not detected an ionized absorber in any of the quasars with intrinsic NALs, even though such features are common in the X-ray spectra of Seyfert galaxies. Contributing factors to this were the low $S/N$ of the X-ray spectra and the high redshift of the quasars themselves (the signature of a warm absorber is most pronounced at energies that are redshifted out of the observable band). Detecting or placing limits on a warm absorbing medium is important because it provides a test of models for the structure of the outflow (especially the ionization structure at high latitudes above the disk; see, for example, Proga et al. 2000). Moreover, a warm medium has been invoked to interpret the variability of mini-BALs (see the discussion of HS1603+3820 by Misawa et al. 2007a). \\item We have not detected high-velocity X-ray absorption lines in any of our spectra. This is not a surprise since such lines are rare. More specifically, unresolved lines have been found in about half a dozen Seyfert galaxies and quasars at $z<0.1$, including MCG-5-23-16 (Braito et al. 2006), IC~4329a (Markowitz et al. 2006), IRAS~13197-1627 (Dadina \\& Cappi 2004), Mrk~509 (Dadina et al. 2005), and PG~1211+143 (Pounds et al. 2007). The observed outflow velocities are of order $0.1\\; c$, while the rest-frame equivalent widths typically range between 10 and 100~eV. Broader, and somewhat stronger absorption lines have also been detected in a handful of high-redshift quasars ($z\\sim 2$--4), namely PG~1115+080 (Chartas et al. 2003, 2007a), H~1413+117 (Chartas et al 2007b), and APM~08279+5255 (Chartas et al. 2002; Hasinger et al. 2002). In these cases the outflow speeds are of order a few tenths of the speed of light and the rest-frame equivalent widths range between a few hundred eV and a few keV. In addition to the cosmological implications of high-velocity X-ray absorption lines noted in \\S\\ref{sec:intro}, their detection will also help us constrain the ionization state of the NAL gas. The limits we have set for Q0014+8118 are rather stringent ($< 28$~eV), comparable to the weakest lines ever detected. In other quasars in our sample the limits are considerably higher, thus uninteresting, owing to the worse $S/N$ in their X-ray spectra. Nevertheless, it is interesting that the absence of high-velocity X-ray absorption lines in Q0014+8118 coincides with the absence of intrinsic UV NALs. \\item Our results, especially Figure~\\ref{fig:correlations}, raise the issue of whether the UV and X-ray properties of NAL quasars connect smoothly to those of BAL quasars. Related to this issue is the question of where do mini-BALs fall in the grand scheme of things. Even though we have found some tantalizing trends, our sample of objects is too small for a firm conclusion on these issues. In this respect, further observations of mini-BALs would be particularly useful because they may fill in the gap between NALs and BALs in Figure~\\ref{fig:correlations}. \\end{itemize} The above questions can be addressed by employing two complementary strategies in future observations. Long exposures of selected objects can yield high-$S/N$ X-ray spectra that can be used for a sensitive search for warm absorbers or high-velocity X-ray NALs. At the same time, (relatively shallow) observations of larger samples of quasars hosting NALs and especially mini-BALs will allow us to explore connections between the corresponding quasar populations." }, "0801/0801.1237_arXiv.txt": { "abstract": "Intrinsic stellar variability can hinder the detection of shallow transits, particularly in space-based data. Therefore, this variability has to be filtered out before running the transit search. Unfortunately, filtering out the low frequency signal of the stellar variability also modifies the transit shape. This results in errors in the measured transit depth and duration used to derive the planet radius, and orbital inclination. We present an evaluation of the magnitude of this effect based on 20 simulated light curves from the CoRoT blind exercise 2 (BT2). We then present an iterative filter which uses the strictly periodic nature of the transits to separate them from other forms of variability, so as to recover the original transit shape before deriving the planet parameters. On average with this filter, we improve the estimation of the transit depth and duration by 15\\% and 10\\% respectively. ", "introduction": "\\subsection{Planet parameters from observations and associated errors} The radius ($R_{\\rm p}$, eq. \\ref{eq1}) and the mass ($M_{\\rm p}$, eq. \\ref{eq2}) of an exoplanet can be fully solved when measuring both the flux and the radial velocity variations of the parent star due to its orbiting planetary companion. \\begin{equation} R_\\textrm{p}=R_{\\star} \\sqrt{\\frac{\\Delta F}{F}} \\label{eq1} \\end{equation} \\begin{equation} M_\\textrm{p}=M_{\\star}^{\\frac{2}{3}}\\frac{K}{\\sin{i}}\\left(\\frac{P}{4\\pi G}\\right)^{\\frac{1}{3}} \\label{eq2} \\end{equation} \\noindent where, $R_{\\star}$ and $M_{\\star}$ are the radius and the mass of the parent star, $\\frac{\\Delta F}{F}$ the flux variation due to the planet transiting the disc of its parent star, $K$ the amplitude of the radial velocity variation of the parent star due the gravitational influence of its orbiting planet, $i$ and $P$ the orbital inclination and period of the planet, and $G$ the gravitational constant. $\\frac{\\Delta F}{F}$, $P$, and $i$ can be measured from the light curve. A common way to measure $R_{\\star}$ and $M_{\\star}$ is by comparing the stellar spectrum to stellar atmosphere models, allowing to derive the stellar parameters ($T_{eff}$, $\\log{g}$, [Fe/H]) used to obtain the stellar mass and radius. $R_{\\star}$ can also be measured more precisely, and without the use of models, with interferometry, or with transit fitting (in the case of high precision light curves). \\subsection{Planet parameters and planet evolution and formation models} Improving the precision on observational planet masses and radii is important for both planet structure and planet formation models. The internal structure of a planet can be studied by comparing its mass and radius to model predictions of planets with different composition. Determining planet structure is important to derive observational statistics on planet types, which can then be compared to the predictions of planet formation models. \\cite{SKHal07} show that to determinate the composition of sub-Uranus planets, error bars of 2\\% on the planet parameters are required. The current uncertainties on planet masses and radii are of the order of 10\\%. Improving these measurements is thus vital to help confirm the models. \\subsection{Sources of uncertainties on planet parameters} The uncertainties on the planet mass, radius, and inclination depend on the uncertainties on the host star mass and radius, on the uncertainties on the transit parameters ($\\frac{\\Delta F}{F}$, total transit duration), and on the uncertainties on the radial velocity measurements. For large planets ($>$Jupiter), the uncertainties on the planet mass and radius are mainly due to the uncertainties on the stellar parameters. For sub-Uranus planets around active stars, the uncertainties on the planet mass and radius can be dominated by the uncertainties on the transit parameters. ", "conclusions": "Based on 20 simulated CoRoT light curves from the BT2 light curve sample, we have shown that \\cite{AI04} pre-detection filter, used to remove stellar variability prior to transit detection, deforms the transit depth by 20\\% and the transit duration by 15\\%. To circumvent this, we have adapted \\cite{KBN05} iterative filtering method to the case of filtering stellar variability present in space-based light curves. The resulting post-detection iterative filter improves the estimation of the transit depth and duration by 15\\% and 10\\% respectively. \\\\ The two areas where we plan to focus future efforts are {\\it a)} to further automate the filtering process (some user interaction is currently needed to initialize the filter smoothing length), and {\\it b)} to evaluate the improvement in the planet parameters uncertainties resulting from the improvement in the transit parameters from this work. As {\\it b)} depends on the particulars of each system, we plan to derive these uncertainties using Monte Carlo simulations for known planets. Further tests will include using other pre-detection filters, including ones which can be applied to data with significant temporal gaps." }, "0801/0801.2287_arXiv.txt": { "abstract": "{ Neuh\\\"auser et al. (2005) presented direct imaging evidence for a sub-stellar companion to the young T~Tauri star GQ Lup. Common proper motion was highly significant, but no orbital motion was detected. Faint luminosity, low gravity, and a late-M/early-L spectral type indicated that the companion is either a planet or a brown dwarf. } { We have monitored GQ Lup and its companion in order to detect orbital and parallactic motion and variability in its brightness. We also search for closer and fainter companions. } { We have taken six more images with the VLT Adaptive Optics instrument NACO from May 2005 to Feb 2007, always with the same calibration binary from Hipparcos for both astrometric and photometric calibration. By adding up all the images taken so far, we search for additional companions. } { The position of GQ Lup A and its companion compared to a nearby non-moving background object varies as expected for parallactic motion by about one pixel ($2 \\cdot \\pi$ with parallax $\\pi$). We could not find evidence for variability of the GQ Lup companion in the K$_{\\rm s}$-band (standard deviation being $\\pm 0.08$ mag), which may be due to large error bars. No additional companions are found with deep imaging. } { There is now exceedingly high significance for common proper motion of GQ Lup A and its companion. In addition, we see for the first time an indication for orbital motion ($\\sim 2$ to 3 mas/yr decrease in separation, but no significant change in the position angle), consistent with a near edge-on or highly eccentric orbit. We measured the parallax for GQ Lup A to be $\\pi = 6.4 \\pm 1.9$ mas (i.e. $156 \\pm 50$ pc) and for the GQ Lup companion to be $7.2 \\pm 2.1$ mas (i.e. $139 \\pm 45$ pc), both consistent with being in the Lupus I cloud and bound to each other. } \\titlerunning{Astrometric and photometric monitoring of the GQ Lup system} ", "introduction": "Based on three epochs of imaging data spanning five years, Neuh\\\"auser et al. (2005, henceforth N05) presented evidence that the $\\le $ few Myr young T Tauri star GQ Lup has a co-moving companion with K$_{\\rm s} \\simeq 13.1$ mag about 0.7$^{\\prime \\prime}$ west ($\\sim 100$ AU at $\\sim 140$ pc) of the primary star. A low-resolution spectrum gave evidence that it has a late-M to early-L spectral type. Temperature and luminosity can be used to estimate the mass via theoretical evolutionary models. They are, however, very uncertain at young ages (up to at least $\\sim 10$ Myr; Chabrier et al. 2005). Using the Wuchterl \\& Tscharnuter (2003) model extended to planetary masses and few Myr of age, the GQ Lup companion could be a 1 to 3 M$_{\\rm Jup}$ object; whereas for the Tucson group models (Burrows et al. 1997), it is few to $\\sim 30$ M$_{\\rm Jup}$; and for the Lyon group models (Baraffe et al. 2002), it is few to $\\sim 40$ M$_{\\rm Jup}$. A comparison with the GAIA-1 model atmospheres indicated low gravity and, hence, a young age and very low mass (N05). Marois et al. (2007) re-analyzed archival HST and Subaru data to study the spectral energy distribution of GQ Lup A and its companion, showing possible evidence for an excess in the L- and R-bands (for the companion), possibly due to a disk and H$\\alpha$ emission; a fit of the data to the dusty GAIA model atmospheres confirmed the temperature and luminosity of the companion given in N05 and revised its radius to $0.38 \\pm 0.05$~R$_{\\odot}$, yielding a mass between 10 and 20 M$_{\\rm Jup}$. McElwain et al. (2007) then obtained higher-resolution spectra (R $\\simeq 2000$) of the GQ Lup companion in the J- and H-bands with the integral field spectrograph OSIRIS at Keck and found a slightly higher temperature (M6-L0) than in N05 (M9-L4), explained by the fact that H$_{2}$ collision-induced absorption is important for low gravity objects according to Kirkpatrick et al. (2006), but not considered in N05. Higher-resolution spectra in the J-, H-, and K-bands taken with VLT/Sinfoni compared to GAIA-2 models could better constrain the parameters of the companion: the temperature is $2650 \\pm 100$ K, the gravity $\\log~g = 3.7 \\pm 0.5$ dex, and the radius $3.5^{+1.50} _{-1.03}$ R$_{\\rm Jup}$ (Seifahrt et al. 2007). Hence, its mass can be as low as few M$_{\\rm Jup}$, but also much higher. Comparing its parameters with 2M0535 A \\& B, where masses have been determined dynamically (an eclipsing double-lined binary brown dwarf in Orion, similar age as GQ Lup, Stassun et al. 2006), could give an upper mass limit of 35 M$_{\\rm Jup}$ (Seifahrt et al. 2007). Thus, the companion to GQ Lup can be regarded as a {\\em planet candidate} according to the best guess value of its mass, which is at or below the brown dwarf desert ($\\sim 30$ M$_{\\rm Jup}$; Grether \\& Lineweaver 2006), proposed as the deviding line between planets and brown dwarfs. Also, it is possible that the true mass is below 13 M$_{\\rm Jup}$, a more conservative upper mass limit for planets. The large error in the luminosity of the companion, which is used for the mass estimate from evolutionary models, is mostly due to a large distance uncertainty, assuming that the object is in Lupus I ($\\pm 50$ pc, N05). Hence, a direct parallax measurement would yield a stronger contraint on luminosity and, hence, mass. To finally confirm that the fainter object near GQ Lup is really a bound companion (rather than, e.g., another member of the Lupus I cloud not orbiting GQ Lup), one would need to see orbital motion. For both measurements, parallax and orbital motion, we have monitored GQ Lup and its companion from 2005 to 2007 by taking six new images with Adaptive Optics. In Sects. 2 \\& 3, we explain the observations, data reduction and astrometric results. We also use the data to monitor the brightness of GQ Lup A and its companion and search for photometric variability (see Sect. 4). We note that Seifahrt et al. (2007) found emission lines in the near-infrared spectra of the companion, indicative of accretion, so some variability is expected. We then add up all imaging data thus far available to obtain a very deep and high-dynamic-range image to search for additional, fainter and/or closer companions (see Sect. 5). We note that both Debes \\& Sigurdsson (2006) and Boss (2006) argued that if the GQ Lup companion is a planet, it should have been moved to its current large separation ($\\sim 100$ AU) from an originally closer orbit by an encounter with another more massive proto-planet, which may be detectable. We use the newly determined parameters to re-estimate some physical parameters of the GQ Lup companion in Sect. 6 and summarize our results at the end. ", "conclusions": "Given the newly constrained mean K$_{\\rm s}$-band magnitude of $13.39 \\pm 0.08$ mag and distance measured for the companion of $139 \\pm 45$ pc (this paper) and with B.C.$_{\\rm K}= 3.0 \\pm 0.1$ mag (following Golimowski et al. 2004) and the temperature of the companion newly constrained in Seifahrt et al. (2007), we can re-estimate the luminosity of the companion to be $\\log (L_{\\rm bol}/L_{\\odot}) = -2.375 \\pm 0.245$, similar to the value in N05, but with a smaller error bar. With the temperature $2650 \\pm 100$~K and gravity $\\log g = 3.7 \\pm 0.5$ dex (g in $g/cm^{2}$) for the GQ Lup companion (Seifahrt et al. 2007), we can use luminosity and temperature to re-calculate its radius to be $3.0 \\pm 0.5$ R$_{\\rm Jup}$. With radius and gravity, we obtain a mass of $\\sim 20$ M$_{\\rm Jup}$ with a possible minimum (value $- 1 \\sigma$) being only few M$_{\\rm Jup}$, and the maximum being around the sub-stellar limit. However, the upper mass limit for the GQ Lup companion is still $36 \\pm 3$~M$_{\\rm Jup}$, because the GQ Lup companion is smaller, cooler, and fainter than both components in the eclipsing double-lined spectroscopic binary brown dwarf 2M0535 (Stassun et al. 2006), which has a similar age as GQ Lup, as already noticed by Seifahrt et al. (2007). We can also use luminosity L, temperature T, gravity g, radius R, and the age of the young T Tauri star GQ Lup ($\\le 2$ Myr, N05, having strong IR excess) to estimate the mass of the companion from theoretical evolutionary models (as done in N05 with the original, less constrained parameters): from Burrows et al. (1997), we consistently obtain for all combinations of L, T, g, R, and age a mass of $\\sim 20$ M$_{\\rm Jup}$, and from Baraffe et al. (2002), we consistently get for all combinations L, T, and age a mass of $\\sim 20$ M$_{\\rm Jup}$. According to the calculations following Wuchterl \\& Tscharnuter (2003), as plotted in Fig. 4 in N05, the GQ Lup companion would have $\\sim 5$ M$_{\\rm Jup}$. It may be seen as intriguing that both the atmospheric and the conventional evolutionary models consistently give $\\sim 20$ M$_{\\rm Jup}$ as the best value. However, we note that the models by Burrows et al. and Baraffe et al. may not be valid for very young objects ($\\le 10$ Myr), as initial conditions are not taken into account, and none of the models used are tested positively for very low-mass objects or calibrated. A better mass estimate can be obtained in the future by comparison with more very young objects with dynamically determined masses and/or atmospheric or evolutionary models that are calibrated. We can summarize our results as follows. \\begin{enumerate} \\item With precise relative astrometry at six new epochs, we have detected small deviations in the separation between GQ Lup A and its companion, consistent with orbital motion of $\\sim 2$ to 3 mas/yr. \\item It remains unclear whether the slight decrease in separation observed is due to orbital motion around each other (bound) or due to different parallaxes (unbound). The latter case is, however, less likely. \\item By comparing the position of GQ Lup A and its companion to a third object in the small NACO FoV, a background star at 800 to 1700 pc, we could determine the parallaxes of GQ Lup A and its companion corresponding to a distance of $156 \\pm 50$ pc for A and $139 \\pm 45$ pc for the companion. \\item Apart from that background star, no additional companion candidates are detected. Outside of 115 mas (18 AU), we can exclude further companion candidates as bright or brighter than the known co-moving companion. \\item We could confirm photometric variability of GQ Lup A in the K$_{\\rm s}$-band (7 exposures over 3 years); variability of the companion is smaller than $\\pm 0.08$ mag. \\item From the newly constrained distance and the mean K$_{\\rm s}$-band magnitude, we re-estimate luminosity, radius, and mass of the companion, but we can still not decide whether the companion is below the deuterium burning mass limit (or below the radial velocity brown dwarf desert, which is at $\\sim 30$ M$_{\\rm Jup}$ according to Grether \\& Lineweaver 2006), i.e., whether the companion is of planetary mass or a brown dwarf. The large uncertainty in mass is due to uncertainties in the theoretical models and the gravity and distance measurements. \\end{enumerate}" }, "0801/0801.2764_arXiv.txt": { "abstract": "We present high-resolution interferometric imaging of LH\\,850.02, the brightest 850- and 1200-\\micron\\ submillimetre (submm) galaxy in the Lockman Hole. Our observations were made at 890\\,\\micron\\ with the Submillimetre Array (SMA). Our high-resolution submm imaging detects LH\\,850.02 at $\\gsim$6$\\sigma$ as a single compact (size $\\lsim$1\\,arcsec or $\\lsim$8\\,kpc) point source and yields its absolute position to $\\sim$0.2-arcsec accuracy. LH\\,850.02 has two alternative radio counterparts within the SCUBA beam (LH850.02N \\& S), both of which are statistically very unlikely to be so close to the SCUBA source position by chance. However, the precise astrometry from the SMA shows that the submm emission arises entirely from LH850.02N, and is {\\it not} associated with LH850.02S (by far the brighter of the two alternative identifications at 24--\\micron\\ ). Fits to the optical-infrared multi-colour photometry of LH850.02N \\& S indicate that both lie at $z\\approx 3.3$, and are therefore likely to be physically associated. At these redshifts, the 24\\micronend--to--submm flux density ratios suggest that LH\\,850.02N has an Arp\\,220-type starburst-dominated far-IR SED, while LH\\,850.02S is more similar to Mrk\\,231, with less dust-enshrouded star-formation activity, but a significant contribution at 24--\\micron\\ (rest-frame ~$5-6$\\, \\micron ) from an active nucleus. This complex mix of star--formation and AGN activity in multi--component sources may be common in the high redshift ultraluminous galaxy population, and highlights the need for precise astrometry from high resolution interferometric imaging for a more complete understanding. ", "introduction": "\\begin{figure*} \\epsfig{figure=lh850-2N.ps,width=80mm} \\epsfig{figure=lh850-2S.ps,width=80mm} \\caption{Optical and near-IR photometry for LH\\,850.02N and LH\\,850.02S, along with fits to their spectral energy distributions (SEDs), and photometric redshifts \\citep{dye2008}. Downward arrows indicate upper limits.} \\label{fig:photoz} \\end{figure*} It has been well established that up to half of the far-infrared (far-IR) extragalactic background is produced by dusty starbursts and active galactic nuclei \\citep{hauser1998,dwek1998,fixsen1998,pei1999}. A significant fraction of this background was resolved at 850\\,\\micron\\ into discrete point sources \\citep{smail1997,hughes1998,barger1998} by the Submm Common-User Bolometric Array \\citep[SCUBA:][]{holland1999} on the 15-m James Clerk Maxwell Telescope (JCMT). These so-called submm galaxies (SMGs) are thought to be high-redshift ultraluminous and hyperluminous IR galaxies \\citep[see][]{chapman2005} that represent massive systems in the process of formation \\citep{scott2002,blain2004}, and may dominate cosmic star formation for nearly the first half of the lifetime of the Universe \\citep[$z\\gsim 1$;][]{blain1999,blain2002}. Since their discovery a decade ago, a series of surveys using SCUBA at 850\\,\\micron\\ \\citep{barger1999,eales1999,eales2000, chapman2002,cowie2002,scott2002,borys2002,webb2003,borys2003,serjeant2003,wang2004,knudsen2006,coppin2006,knudsen2008} and mm instruments at longer wavelengths \\citep{greve2004,dannerbauer2004, laurent2005,dscott2006,bertoldi2007,scott2008} have amassed catalogues of hundreds of SMGs. Unfortunately, the detailed study of these objects has been hindered somewhat by the poor angular resolution ($\\sim$11--18\\,arcsec) of current submm/mm telescopes. This problem was first addressed via deep radio continuum surveys, which exploited the radio--FIR correlation \\citep[see][for a review]{condon1992} in combination with statistical arguments \\citep[e.g.,][]{ivison2002, ivison2007} to associate 1.4-GHz sources with the submm emission. Radio counterparts allowed SMGs to be localised with sub-arcsec precision and hence allowed a more detailed study of their properties. Optical spectroscopy of these radio-identified samples confirmed that SMGs lie preferentially at high redshift \\citep[median $z\\sim 2.3$;][]{chapman2005} and enabled CO spectroscopy which revealed them to be compact, massive, gas-rich, possibly merging systems \\citep[e.g.,][]{neri2003,sheth2004,kneib2005,greve2005,tacconi2006}. Despite the undoubted success of deep radio imaging of SMGs, there is still a clear need for high-resolution {\\it submillimetre/millimetre} observations of at least a subset of the SMGs found in current, complete samples. Specifically, two classes of object require such observations to locate the source of the submillimetre emission with the required astronomical precision for optical/infrared follow-up. First, because of the rapid dimming of the radio continuum with redshift \\citep[$I_\\nu\\sim (1+z)^{\\alpha-3}$, $\\alpha=-0.8$, where $I_{\\nu}\\propto\\nu^{\\alpha}$;][]{condon1992}, even the deepest existing radio imaging is relatively insensitive to SMGs at $z\\gsim 3$; as a result, typically only around two thirds of SMGs have been detected at radio wavelengths \\citep{ivison2002}. Other techniques \\citep{ashby2006,pope2006} have been suggested, which make use of data from the IR Array Camera \\citep[IRAC:][]{fazio2004}, in combination with 24-\\micron\\ observations using the Multiband Imaging Photometer \\citep[MIPS:][]{rieke2004}, on board the {\\it Spitzer Space Telescope} to select counterparts. However, they too may have hidden biases which are difficult to quantify, and the only unambiguous way to select the correct optical counterparts for these radio-unidentified SMGs is via time-intensive submm/mm high-resolution interferometric imaging. Interferometric observations at mm \\citep{downes1999,frayer2000,dannerbauer2002,genzel2003, greve2005,tacconi2006,dannerbauer2008} and submm \\citep[][see also Iono et al. 2008, in prep.]{iono2006,wang2007} wavelengths have successfully detected a growing catalogue of SMGs (in the process also confirming the reliability of the radio--submm association where a unambiguous radio counterpart has already been discovered). Most recently, \\citet{younger2007} followed up a flux-limited sample of seven mm-selected SMGs detected by the AzTEC camera \\citep{wilson2008} at the JCMT -- including five without reliable radio identifications -- at 890\\,\\micron\\ with the Submillimeter Array \\citep[SMA:][]{ho2004}, the counterparts of which suggested a population of very luminous SMGs at higher redshift than radio-identified samples (see also Yun et al. 2008, in prep.). That this bright (median 890-\\micron\\ flux density 12.0\\,mJy), AzTEC-selected sample contains a significant high-redshift tail of SMGs supports earlier evidence that the brightest SMGs may be the most distant \\citep[e.g.\\ Figure~9 of][]{ivison2002,dunlop2001}. Second, within the radio-identified sub-samples of SMGs there remains some confusion. Specifically, there exists a statistically significant fraction of SMGs which possess more than one statistically robust radio counterpart \\citep[$\\sim20\\pm5\\%$;][]{ivison2007}. Monte--carlo simulations suggest that these associations are observed significantly more frequently than would be expected from chance associations. The nature of these multiply-identified SMGs remains somewhat uncertain, although the steepness of the SMG number counts suggests that the bright submillimetre sources are only rarely expected to arise from the blending/confusion of more moderate luminosity subcomponents. Interferometric imaging of the rest--frame far--IR continuum offers the best way to investigate the true nature of these interesting systems by precisely locating the source of the submm emission \\citep[see e.g., SMM J094303+4700 H6 and H7;][]{tacconi2006}. In this paper, we present high-resolution SMA 890-\\micron\\ interferometric imaging of LH\\,850.02/LH\\,1200.04 (hereafter simply LH\\,850.02), the brightest source in both the MAMBO 1200-\\micron\\ \\citep{greve2004} and SCUBA 850-\\micron\\ \\citep{coppin2006} maps of the Lockman Hole. In \\S~\\ref{sec:obs} we describe our observations and data reduction, and in \\S~\\ref{sec:discuss} we discuss some possible implications of our results. Throughout this paper, we use the Vega magnitude system, and assume a flat concordance cosmology with $\\rm (\\Omega_m,\\Omega_\\Lambda,H_0) = (0.3,0.7,70$ km\\,s$^{-1}$\\,Mpc$^{-1}$). ", "conclusions": "We present high-resolution 890-\\micron\\ interferometric imaging of LH\\,850.02, a bright SMG in the LH, with the SMA. We detect LH\\,850.02 at $\\gsim 6\\sigma$ as a single compact (size $\\lsim$1\\,arcsec, or $\\lsim$8\\,kpc) point source, and determine a position accurate to $\\sim$0.2\\,arcsec. From this, we identify multi-wavelength counterparts and find that only one (LH850.02N) of two candidate radio counterparts is associated with the submm emission. The nearby radio source with strong 24-\\micron\\ emission (LH850.02S) does {\\em not} contribute significantly to the submm flux. In this case, and by analogy other SMGs with multiple radio counterparts, the radio continuum more reliably locates the source of the submm -- and with it the properties of the associated starburst -- than does the mid--IR; similar to the results of \\citet{younger2007}. Since both radio sources have similar photometric redshifts, and therefore may be physically associated, their respective 24\\micronend--to--submm suggest that LH\\,850.02N has an Arp\\,220-type starburst-dominated far-IR SED, while LH\\,850.02S has a Mrk\\,231-type SED with a significant mid-infrared contribution from an active nucleus. As a result of the relatively shallow, wide-field surveys conducted to date with the AzTEC camera -- wherein objects of this type were first found in significant numbers \\citep{younger2007} -- existing 1100-\\micron\\ samples are typically $\\gsim$2$\\times$ brighter than samples selected with SCUBA or MAMBO. We suggest that the recent prevalence of candidate high-redshift SMGs is more closely related to their high flux densities than to the long survey wavelength, a tendency noted by \\citet{ivison2002} and \\citet{wall2008}. It is therefore perhaps not surprising that we have here constrained the brightest source in the SHADES survey of the LH, LH\\,850.02, to lie at high redshift $z\\gsim 3$." }, "0801/0801.0557_arXiv.txt": { "abstract": "I discuss the relationship between mass loss and nucleosynthesis on the Asymptotic Giant Branch (AGB). Because of thermal pulses and possibly other mixing processes, products of nucleosynthesis can be brought to the surface of AGB stars, increasingly so as the star becomes more luminous, cooler, and unstable against pulsation of its tenuous mantle. As a result, mass loss is at its most extreme when dredge-up is too. As the high rate of mass loss truncates AGB evolution, it determines the enrichment of interstellar space with the AGB nucleosynthesis products. The changing composition of the stellar atmosphere also affects the mass-loss process, most obviously in the formation of dust grains --- which play an important r\\^ole in driving the wind of AGB stars. ", "introduction": "Asymptotic Giant Branch (AGB) stars span a range in masses from $<1$ M$_\\odot$ to $\\sim8$ M$_\\odot$, which makes them important diagnostics of the star formation history and important players in galactic chemical enrichment from ages as young as $<100$ Myr up to (nearly) a Hubble time. They shed much, often most, of their initial mass in the form of dusty winds before leaving a white dwarf behind. In this review I will first ask the question why we care about mass loss, with a bias towards those amongst us who are interested in nucleosynthesis. I then briefly discuss what determines the mass-loss rate, and comment on ways in which nucleosynthesis may, or may not, alter the conditions for the mass-loss mechanism. ", "conclusions": "" }, "0801/0801.0885_arXiv.txt": { "abstract": "Model of supermassive black holes formation inside the clusters of primordial black holes is developed. Namely, it is supposed, that some mass fraction of the universe $\\sim10^{-3}$ is composed of the compact clusters of primordial (relic) black holes, produced during phase transitions in the early universe. These clusters are the centers of dark matter condensation. We model the formation of protogalaxies with masses about $2\\,10^{8}{\\rm M_{\\odot}}$ at the redshift $z=15$. These induced protogalaxies contain central black holes with mass $\\sim10^5{\\rm M_{\\odot}}$ and look like dwarf spheroidal galaxies with central density spike. The subsequent merging of induced protogalaxies and ordinary dark matter haloes corresponds to the standard hierarchical clustering scenario of large-scale structure formation. The coalescence of primordial black holes results in formation of supermassive black holes in the galactic centers. As a result, the observed correlation between the masses of central black holes and velocity dispersion in the galactic bulges is reproduced. ", "introduction": "The problem of galaxy formation with supermassive central black hole (BH) becomes more and more intriguing and ambiguous in view of discovery of distant quasars at redshifts $z>6$ in Sloan Digital Sky Survey \\cite{z6}. The maximum observed red-shift $z=6.41$ belongs to the quasar with luminosity corresponding to the accretion onto BH with the mass $3\\,10^{9}{\\rm M_{\\odot}}$ \\cite{will03}. Such an early formation of BHs with masses $\\sim10^{9}{\\rm M_{\\odot}}$ is a serious difficulty for the standard astrophysical models of supermassive BH formation in galaxies supposing a fast dynamical evolution of the central stellar clusters in the galactic nuclei (see e.~.g. \\cite{spitzer,saslaw,lightshap78,rees84,dokrev} and references therein), a gravitational collapse of supermassive stars and massive gaseous disks in galactic centers (see e.~.g. \\cite{rees84,bintrem87,el2,haehreeesnar98}), the multiple coalescences of stellar mass BHs in galaxies (see e.~.g. \\cite{rees92,MouTan02,GebRicHo02,kawaguchi}) with the subsequent multiple merging of galactic nuclei in collisions of galaxies in clusters (see e.~g. \\cite{ValVal89,cattaneo05,komossa03,smirnova06,springel05,escala05,masjedi06}). All standard astrophysical (or galactic) scenarios of supermassive BH origin predict a rather late time of supermassive BH formation in the galactic nuclei. An other difficulty is that all these astrophysical scenarios are realized only in strongly evolved galactic nuclei. In view of these problems the cosmological scenarios of massive primordial BHs formation become attractive \\cite{zeld67,carr75,DolgovSilk93,quin,Ru1,carr05}. In cosmological scenarios the seeds of supermassive BHs are formed long before the formation of galaxies. These primordial black holes (PBH) can be the centers of baryonic \\cite{Ryan} and dark matter (DM) \\cite{DokEroPAZH} condensation into the growing protogalaxies. There are proposed two alternative possibilities: (i) a formation of initially massive primordial BHs and their successive growth up to $\\sim10^9{\\rm M_{\\odot}}$ due to accretion of ambient matter or (ii) a formation of small-mass primordial BHs and their subsequent merging into the supermassive ones in the process of hierarchical clustering of protogalaxies. An effective cosmological mechanism of massive primordial BH formation and their clusteri\\-zation was developed in works \\cite{Ru1,Ru2,KR04}. In this papers the properties of spherically symmetric primordial BH clusters were investigated. As a basic example, a scalar field with the tilted Mexican hat potential had been accepted. The properties of resulting primordial BH clusters appear to be strongly dependent on the value of initial phase. In addition, the properties of these clusters depend on the tilt value of the potential $\\Lambda $ and the scale of symmetry breaking $f$ at the beginning of inflation stage. As a result, the mass distribution of primordial BH clusters could vary in a wide range. In our previous paper \\cite{weqso} we considered the model parameters leading initially to large clusters with a rather heavy mass of the central primordial BH, $\\sim4\\,10^7{\\rm M_{\\odot}}$. These central heavy primordial BH can grow due to accretion up to $\\sim10^{9}{\\rm M_{\\odot}}$ and, therefore, may explain the observed early quasar activity. The elaboration of a discussed mechanism of cosmological primordial BH formation is far from completion and detailed elaboration. For example, there are no now any physical substantiations for reconstruction of scalar field potential parameters and initial characteristics of primordial BH clusters. It is connected not only with the uncertainties of observational data but also with complexities of phase transition details. For example, the domain walls formed during the phase transition in the early universe has a topology of sphere but with a very complicated surface form. When these closed domain walls are turned out inside the horizon, they become self-gravitating. Inside the horizon domain walls tend to obtain a spherical form due to surface tension, but at the same time they strongly oscillate and generate gravitational and scalar waves. As a result their mass gradually diminish. An approximate consideration of this effect \\cite{Ru2} demonstrate that for a wide range of initial theoretical free parameters there are conditions for formation of cluster with a supermassive primordial BH. For this reason we will not fix here the definite values of free parameters in the discussed cosmological model of massive black hole formation. The influence of non-sphericity of the formed domain walls see in \\cite{Ru2,KR04}. An application of the mechanism found in \\cite{Ru1,Ru2,KR04} is not limited by a specific form of the scalar field potential. Below we demonstrate that substantial number of potentials, like e.~g. those used in hybrid inflation, also result in formation of massive BHs. Moreover, it is hard to avoid primordial BHs overproduction in the early Universe. In fact, any inflationary model using potential with two or more minima must take into account this mechanism of primordial BHs overproduction. In this paper we choose parameters of the potential which lead to formation of relatively small primordial BH clusters. We suppose that relatively small primordial BH clusters provide the major contribution to initial density pertur\\-bations which afterwards evolve into protogalaxies. The hierarchical clustering of protogalaxies during the cosmological time leads to the observable large scale structure. We describe the gravitational dynamics of DM coupled with primordial BH clusters and demonstrate that a protogalaxy could be formed without any initial fluctuations in DM density. In this case the clusters of primordial BHs play the role of initial fluctuations. Two scenarios of supermassive BHs formation could coexist: (i) the most massive clusters of primordial BHs account for an early quasar activity \\cite{weqso}, but (ii) less massive (considered in this paper) clusters of primordial BHs produce more numerous supermassive BHs observed nowadays in almost all structured galaxies. There are several stages of BHs and galaxies formation in the described scenario: (i) Formation of closed walls of scalar field just after the end of inflation with a subsequent collapse some of these walls with formation of massive primordial BH cluster with the most massive BH in the center after the horizon crossing according to \\cite{Ru1,Ru2}. (ii) Detachment of the central dense region of the primordial BH cluster from cosmological expansion and virialization. Numerous small-mass BHs merge with a central one. (iii) Detachment of the outer cluster region (where DM particles dominate) from cosmological expansion and a protogalaxy growth. Termination of a protogalaxy growth due to interaction with the surrounding standard DM fluctuations. (iv) Gas cooling and star formation accompanied by the merging of protogalaxies and final formation of modern galaxies. ", "conclusions": "We describe here a new model of protogalaxy formation with the cluster of primordial BHs as a source of initial density perturbation. The used mechanism of primordial BH formation \\cite{KR04,Ru2a} provide us with a set of primordial BH clusters of different total mass. This variety of initial conditions leads, therefore, to the variety of protogalaxies from the very beginning of their formation. In this paper we choose for numerical modeling only those BH clusters which produce the large number of small relatively protogalaxies. This model predicts the very early galaxy and quasar formation. An other inevitable consequence of this model it the existence of intermediate mass BHs beyond the dynamical centers of galaxies and in the intergalactic medium. May be one of these type intermediate mass BHs was already observed by the X-ray Chandra telescope in the galaxy M82 \\cite{Kaar00}. More definitely in this model the protogalaxies are formed at redshift $z=15$. These induced protogalaxies have initially the following parameters: a constituent total mass of DM $M_{\\rm DM}=2.2\\,10^{8}{\\rm M_{\\odot}}$, a virial radius $1.8$~kpc, a mass of central BH $M_{\\mathrm{BH}}=7.2\\,10^{4}M_{\\odot}$. In the following cosmological and dynamical evolution, these protogalaxies are assembled by hierarchical clustering into the nowadays galaxies. The clustering process occurs in a stochastic manner and leads to the specific correlation between the central supermassive BH mass and galactic bulge velocity dispersion \\cite{DokEroPAZH}. An alternative proposed scenario is based on the initial large primordial BH clusters, when a resulting galaxy contains a single primordial BH growing due to accretion of ambient gas and stars and producing early quasar activity \\cite{weqso}. It is worth to estimate in the framework of our model a probability to find a nowadays galaxy without supermassive BH. Induced galaxies (with a central cluster of primordial BHs) and ordinary small protogalaxies have mass $M_{\\mathrm{DM}}=10^{8}M_{\\odot}$, while the modern galaxies are much more massive, $M_{\\rm DM}=10^{12}{\\rm M}_{\\odot}$. The merging of induced galaxy with an ordinary protogalaxy produces a next generation protogalaxy with the massive central BH. Therefore, about $10^{4}$ collisions is required to form the nowadays galaxy. Suppose that an amount of induced galaxies is about 0.1\\% comparing with the ordinary ones. A corresponding probability to find a modern galaxy without supermassive BH is less than $0.999^{10000}\\simeq 4.5\\,10^{-5}$. Hence, even a very small fraction of induced galaxies is able to explain the observable abundance of AGN." }, "0801/0801.3886_arXiv.txt": { "abstract": "{ This report is based on a rapporteur talk presented at the 30th International Cosmic Ray Conference held in Merida, Mexico (July 2007), and covers three of the OG sessions devoted to neutrino, gravitational wave, and $\\gamma$-ray detection.} \\email{growell@physics.adelaide.edu.au} \\begin{document} ", "introduction": "Summarised here are key results from papers and posters presented in the sessions OG~2.5 (neutrino detection), 2.6 (gravitational wave detection), and 2.7 ($\\gamma$-ray detection). The number of presentations in each session was 75 (OG~2.7), 19 (OG~2.5) and 2 (OG~2.6), with $\\gamma$-ray detection clearly dominating. OG~2.5 was devoted to neutrino/$\\gamma$-ray connections and related experimental and theoretical issues and contains overlaps with the HE~2 session. A more detailed summary of neutrino detectors and related astrophysical theory can be found in the HE~2 rapporteur by Tom Gaisser \\cite{Gaisser-RAPP}. Here, summaries of each sessions are ordered according to the number of contributions. I have kept to citing only ICRC presentations/posters since in most cases a detailed list of references may be found therein. ", "conclusions": "Some key conclusions on these sessions primarily devoted to the detection of $\\gamma$-rays and neutrino can be outlined as follows: \\begin{itemize} \\item Ground-based $\\gamma$-ray astronomy is a now a mature field employing two established and complementary techniques (a) (Stereoscopic) Imaging Atmospheric Cherenkov Imaging; (b) Water Chereknov detection of air shower particles. All of the planned and proposed detectors appear to be making use of at least one of these two techniques. \\item The next step beyond H.E.S.S., VERITAS, MAGIC-II etc.. is taking shape via continental-wide organisation efforts such as CTA (Europe) and the WhitePaper (USA), along with several other proposals. This is necessary in order to gather community-wide support for the required funding scales of order 100 Million dollar/Euros. \\item GLAST is ready to go and its launch date is not far away. GLAST will no doubt provide a fresh and clearer look at the MeV to GeV Universe. \\item Neutrino event rates in forthcoming detectors (IceCube, KM3NeT) appear to be a few per year from the strongest $\\gamma$-ray sources. This would give them a chance to realise discovery of extraterrestrial neutrino sources. In addition the first efforts in coordinated neutrino/$\\gamma$-ray observations have begun. \\end{itemize} I wish to thank the organisers of the 30th ICRC for giving me the opportunity to summarise these topics. I also thank in particular Simon Swordy and Tom Gaisser for presenting my slides. \\scriptsize" }, "0801/0801.1551_arXiv.txt": { "abstract": "% We have studied the kinematics of the ionized gas and stellar component in Mrk334 using methods of panoramic (3D) spectroscopy. The observations were performed at the prime focus of the SAO RAS 6-m telescope with the integral-field spectrograph MPFS \\citep{mpfs} and with a scanning Fabry-Perot interferometer (FPI), installed on the multimode device SCORPIO \\citep{afan}. Based on these data, the monochromatic maps and velocity fields in different emission lines of the ionized gas were constructed. The diagnostic diagrams have been made based on the emission lines ratios. ", "introduction": "Mrk334 is a Sy1.8 galaxy, located at the distance of 88 Mpc (the scale is 430 pc/$''$). This galaxy demonstrates merging features as tidal tail and bright region `A' on the west from the nucleus. According \\citet{gonzadelga97}, the region \"A' can be a second nucleus or a remainder of a satellite galaxy, which was devoured by Mrk334. On the MPFS maps in different emission lines, three bright regions can be distinguish: nucleus, regions `A' and `B' (Fig.~\\ref{smir:fig1}). Spectra from these regions differ from each other and the gas ionization state vary from one region to another dramatically: on the diagnostic diagrams, the points belonging to the knot `B' lie in the region of ionization by a nonthermal radiation. All points that correspond to the knot `A' fall into the photoionization region. Surprising, that the nucleus is located in the HII/LINER boundary i.e., the gas can also be partially ionized here by both shocks and radiation from young hot stars, but not by non-thermal source! Also, on the [SII] ratio map the region with the lowest density is the knot `B', whereas the density in the knot `A' is much higher. \\begin{figure} \\plotone{smirnova1.eps} \\caption{ MFPS flux map (left) and velocity field (middle) in the [OIII]$\\lambda$ 5007\\AA\\, emission line. The large-scale residual velocity field in the H$\\alpha$ line, isophotes in this line are overlapped (right).} \\label{smir:fig1} % \\end{figure} ", "conclusions": "" }, "0801/0801.4142_arXiv.txt": { "abstract": "We explore the noncommutative effect on single field inflation and compare with WMAP five-year data. First, we calculate the noncommutative effect from the potential and dynamical terms, and construct the general form of modified power spectrum. Second, we consider the leading order modification of slow-roll, DBI and K-inflation and unite the modification, which means the modification is nearly model independent at this level. Finally, comparing with the WMAP5 data, we find that the modified can be well realized as the origin of the relative large spectral index and the quite small running. ", "introduction": "Inflation \\cite{inflation} is a very successful paradigm for naturally understanding the puzzling aspects of hot big bang cosmology such as flatness, homogeneity and monople problems. As it causally explains the origin of the super-horizon density perturbations, inflation seeds the large structure of the universe we observe today. More like a paradigm than a theory because of the missing fundamental basis, inflation is well established by the observation of cosmic microwave background. Models based on pure de sitter spacetime predict a scale-invariant and adiabatic power spectrum, which can be characterized by $n_s-1=0$. However, this is not the history of our unverse. WMAP and other astrophysical observations predict nontrivial spectral index and running. The WMAP team has reported the five-year data\\cite{WMAP5yr} and used them to constrain the physics of inflation. For the $\\Lambda $CDM model, WMAP5 data show that the index of power spectrum satisfies $n_{s}=0.963_{-0.015-0.028}^{+0.014+0.029}~~~(1\\sigma, 2\\sigma~ \\text{CL}); \\label{nsnumber1}$. Combining WMAP with SDSS and SNIa, the result is $n_{s}=0.960_{-0.013-0.027}^{+0.014+0.026}~~~(1\\sigma, 2\\sigma~ \\text{CL})$. Though the red power spectrum is still favored at the level of $3\\sigma $ CL, the result is a little bluer than that of WMAP3. And, the running of the spectral index is not favored anymore. With WMAP5 data only, the running is $\\alpha _{s}=\\frac{dn_{s}}{d\\ln k}=-0.037_{-0.028}^{+0.028}~~~(1\\sigma~\\text{CL})$. Combining with SDSS and SNIa data, the result is $\\alpha _{s}=\\frac{dn_{s}}{d\\ln k}=-0.032_{-0.020}^{+0.021}~~~(1\\sigma~\\text{CL})$. We find that the spectral index in WMAP5 is slightly larger than that in WMAP3 and the running is much smaller than that in WMAP3. Theoretically, we can modify the inflation models, but it will increase the implicity from the aspect of a fundamental theory. In the series works of \\cite{ brandenberger1, easther1, kaloper}, running index emerges from non-Bunch-Davis vacuum and the new physics will imprint on CMB. However, the serious problem is we still know little about de sitter spacetime, which is noted as the ambiguity of vacuum selection. In the papers \\cite{bean}, another method has been presented. The authors obtained the nontrivial spectral index and its running from the generalized slow-roll inflation with arbitrary sound speed. However, as a new challenge for most inflation models, we expect the spectral index and its running can provide the opportunity to observe the short distance physics and the first moment scenario of our universe. As a candidate to probe the quantum gravity in string theory, noncommutativity is applied to describe D-brane physics. Considering the string effect and strong string interaction, short distance geometry becomes very different. In addition, while the usual concept of geometry is completely break down near the singularity geometry, noncommutativity is also a good description. In the works of \\cite{brandenberger2, huang}, the authors introduced noncommutativity, which naturally emerges from string theory, to inflation physics. As the running of spectral index is large, they obtained nice results with WMAP3 experiment data. However, most recent, WMAP5 explores a quite small running index. In this lecture, we use another method to calculate the modification. We directly calculate the modification of power spectrum from the noncommutative potential and dynamical terms while the quantum modes solution still remains classical. We study the modification of the single field inflation models including slow-roll, DBI and K-inflation. As our calculation includes all the potential and dynamics terms, the modification is exact. We find these three modification can be written in one unite form at leading order. This means that the modification can not be tested by different models. In addition, the leading order of modified power spectrum is $1/k^{2}$, which means that the modification is small. Our calculation shows that the noncommutative potential and dynamical terms can be realized as the possible origin of the spectral index and its running since it is favored by WMAP5 data. This paper is organized as follows. In section 2, we review the inflation formalism with a general lagrangian. In section 3, we study the noncommutative effect on inflation including the modification of potential and dynamic and reconstruct the power spectrum. In section 4, we study concrete models and compare with the WMAP5 observation. Section 5 contains some conclusion and discussion. ", "conclusions": "As an important method to detect short distance physics, noncommutativity naturally emerges from string theory and is applied to inflation physics. Branderberger and Ho found noncommutativity can modify the power spectrum in a significant way. In addition, in papers\\cite{cai}, it has been used to regulate the eternal inflation. Recent released WMAP5 data favor a red-tilted power spectrum of primordial fluctuations at the level of two standard deviations, which is the same as the WMAP3 result. However, qualitatively, the spectral index is slightly greater and the running is quite small than the three-year value. Our calculation is quite different. First, we calculated the modification from the potential and dynamical terms while the solution of inflaton remains classic. Second, we calculated the modification of slow-roll, DBI and K-inflation. We found the modification is model independent, as we only considering the leading order. In addition, as the leading order modification of power spectrum is proportional to $1/k^{2}$, the modification of is quite small and can be favored by WMAP5. Third, as we used the string scale in our paper and the correction is proportional to $(\\theta^{12})^{2}$, it could be larger if we have a relative low noncommutative scale. Finally, as WMAP five-year data provide stringent limits on deviation from the minimal, 6-parameter $\\Lambda $CDM model, we expect more years WMAP and Planck data give us more information about the anisotropy of CMB. In addition, future CMB missions such as ground-based polarization experiment of BICEP and ESA's Planck Surveyor, are also expected to help us faithfully understand the early universe." }, "0801/0801.1767_arXiv.txt": { "abstract": "{We exploit a set of high signal-to-noise ($\\sim 70$), low-resolution ($R\\sim 800$) quasar spectra to search for the signature of the so-called proximity effect in the \\ion{H}{i} Ly$\\alpha$ forest. Our sample consists of 17 bright quasars in the redshift range $2.7 < z < 4.1$. Analysing the spectra with the flux transmission technique, we detect the proximity effect in the sample at high significance. We use this to estimate the average intensity of the metagalactic UV background, assuming it to be constant over this redshift range. We obtain a value of $J = (9 \\pm 4) \\times 10^{-22}$ erg cm$^{-2}$ s$^{-1}$ Hz$^{-1}$ sr$^{-1}$, in good agreement with previous measurements at similar $z$. We then apply the same procedure to individual lines of sight, finding a significant deficit in the effective optical depth close to the emission redshift in every single object except one (which by a different line of evidence does nevertheless show a noticeable proximity effect). Thus, we clearly see the proximity effect as a universal phenomenon associated with individual quasars. Using extensive Monte-Carlo simulations to quantify the error budget, we assess the expected statistical scatter in the strength of the proximity effect due to shot noise (cosmic variance). The observed scatter is larger than the predicted one, so that additional sources of scatter are required. We rule out a dispersion of spectral slopes as a significant contributor. Possible effects are long time-scale variability of the quasars and/or gravitational clustering of Ly$\\alpha$ forest lines. We speculate on the possibility of using the proximity effect as a tool to constrain individual quasar ages, finding that ages between $\\sim 10^{6}$ and $\\sim 10^{8}$ yrs might produce a characteristic signature in the optical depth profile towards the QSO. We identify one possible candidate for this effect in our sample.} ", "introduction": "\\begin{figure*} \\sidecaption \\includegraphics*[width=12cm]{7088fig1.eps} \\caption{Example of a quasar spectrum taken from the dataset (HE 0940$-$1050, $z=3.088$). In the upper panel the flux density in units of erg cm$^{-2}$ s$^{-1}$ \\AA$^{-1}$ as function of wavelength is presented with the uncorrected (dotted line), corrected (long-dashed line) continuum level and the noise of the spectrum (dashed-dotted line). In the lower panel the fitted transmission is shown in gray while the corrected continuum is plotted in black (See Sect.~\\ref{datared} and \\ref{syserrors} for details). In both panels the vertical dotted line represent the emission redshift. } \\label{spec} \\end{figure*} The multitude of absorption lines seen in the spectra of high redshift quasars gives important information about the state of matter in the universe, tracing the physical conditions of the intergalactic medium (IGM) at various epochs. It is commonly believed that for column densities up to $N_{\\mbox{\\tiny\\ion{H}{i}}} \\approx 10^{17.5}\\,\\mathrm{cm^{-2}}$, the absorbers are in photoionisation equilibrium with a metagalactic ultraviolet background field (UVB), composed of the integral over all sources of UV radiation (essentially, star-forming galaxies and quasars). High-resolution spectra of the Lyman forest provide not only a rather detailed statistical characterisation of the absorber properties such as line number densities as well as temperature and density distributions \\citep[e.g.,][]{kim01}, but also physical parameters such as \\ion{H}{i} and \\ion{He}{ii} photoionisation rates \\citep{rauch97,fardal98}, which directly relate to the intensity of the UVB, as described recently by \\citet{bolton05} invoking hydrodynamical simulations. Independently, the UVB has been successfully synthesised by combining the observed quasar luminosity function and the UV emission from galaxies (although the latter is still very uncertain) with the propagation of diffuse radiation \\citep{haardt96}. In the vicinity of strong UV sources such as bright quasars, the \\ion{H}{i} photoionisation rate should locally increase, further reducing the density of residual neutral hydrogen. This enhances the transparency of the IGM to \\ion{H}{i} ionising radiation and should become observable as a weakening of the Lyman forest absorption near such sources. Such an effect has first been noted by \\citet{carswell82} and was later baptised `Inverse' \\citep{murdoch86} or `Proximity Effect' \\citet[][hereafter BDO88]{bajtlik88}. Its prime application has been so far the possibility to derive an independent estimate of the UVB intensity, by measuring the reduction of column densities (against the global evolution of absorption line density increasing with redshift) and combining this with the QSO luminosity at the Lyman limit, assumed to be known. The best constraints on the UVB using the proximity effect stem from the combined analysis of large quasar samples \\citep{cooke97,scott00,liske01}, yielding mostly values consistent with the above quoted other methods. However, the uncertainties are still substantial. Besides the problem that a limited number of lines of sight always suffers from `cosmic variance', there may also be systematic biases. In particular, if QSOs reside in intrinsically overdense environments then the signature of the proximity effect will appear weaker than predicted \\citep{loeb95,rollinde05}. Another uncertainty is the possibly limited lifetime of quasars. On the other hand, the proximity effect may also be used to derive constraints on this important, but largely unknown astrophysical quantity \\citep[e.g.,][]{pentericci02}. In this paper we present an exploitation of new observational material in terms of the proximity effect (Sect.~\\ref{obs_set}). Rather than the traditional line counting we use the more sensitive flux transmission statistic to search for proximity effect signatures, augmented by extensive Monte-Carlo simulations to calibrate the systematic and statistical errors (Sect.~\\ref{lyaforest}). We deliver our results in Sections \\ref{combined} and \\ref{indiv_qsos}. Firstly we briefly present an analysis of the combined sample of 17 QSO spectra and derive an estimate of the UV background intensity. Secondly we demonstrate that the effect is measurable on single sightlines \\citep{williger94,lu96,savaglio97}, not only statistically in large samples \\citep{bajtlik88, scott00, liske01}. The effect can actually be systematically detected in each single QSO of our sample (with one special case that is discussed separately). We speculate about the possibilities to detect signatures of finite quasar ages by virtue of the proximity effect. Throughout this paper, we assume a flat $\\Lambda-$Universe with $\\Omega_\\mathrm{m}=0.3$ and $\\Omega_\\mathrm{\\Lambda}=0.7$ and $H_0=71\\,\\mathrm{km\\,s^{-1}\\,Mpc^{-1}}$. \\begin{table*}[t] \\small\\centering \\caption[]{Log of observations.} \\label{Tab1} \\begin{tabular}{lcccccclc} \\hline\\hline\\noalign{\\smallskip} QSO & $V$ mag &$z_\\mathrm{em}$ &Exp. Time&Seeing & Airmass &Sky Condition$^\\mathrm{a}$ & Obs. Date&Ref. Mag\\\\ &&&(s)&(arcsec)&&& \\\\ \\noalign{\\smallskip}\\hline\\noalign{\\smallskip} \\object{CTQ 0247} & 17.4 & 3.025 & 750 & 1.29 & 1.07 & CL, WI & Dec. 8, 2002 & 4\\\\ \\object{CTQ 1005} & 18.4 & 3.205 & 1500 & 1.13 & 1.21 & TN & Jan. 9, 2003 & 3\\\\ \\object{CTQ 0460} & 17.5 & 3.139 & 900 & 1.45 & 1.10 & PH & Dec. 23, 2002 & 4\\\\ \\object{H 0055$-$2659} & 17.5 & 3.665 & 600 & 1.34 & 1.02 & CL, WI & Dec. 8, 2002 & 4\\\\ \\object{HE 0940$-$1050} & 16.4 & 3.086 & 600 & 0.86 & 1.12 & TN, TK & Nov. 26, 2002 & 3\\\\ \\object{HE 2243$-$6031} & 16.4 & 3.010 & 600 & 1.02 & 1.24 & CL, PH & Nov. 9, 2002 & 4\\\\ \\object{HE 2347$-$4342} & 16.7 & 2.885 & 600 & 0.77 & 1.06 & PH & Nov. 10, 2002 & 1\\\\ \\object{PKS 2126$-$15} & 17.0 & 3.285 & 600 & 0.97 & 1.20 & PH & Oct. 29, 2002 & 4\\\\ \\object{Q 0000$-$26} & 18.0 & 4.098 & 600 & 1.31 & 1.03 & TN, CL & Nov. 7, 2002 & 4\\\\ \\object{Q 0002$-$422} & 17.2 & 2.767 & 600 & 1.39 & 1.45 & TK, TN & Oct. 14, 2002 & 3\\\\ \\object{Q 0347$-$383} & 17.7 & 3.220 & 800 & 1.45 & 1.15 & TN & Jan. 9, 2003 & 1\\\\ \\object{Q 0420$-$388} & 16.9 & 3.120 & 600 & 1.37 & 1.06 & CL, WI & Dec. 8, 2002 & 1\\\\ \\object{Q 0913$+$0715} & 17.8 & 2.787 & 800 & 1.45 & 1.18 & TN & Jan. 9, 2003 & 2\\\\ \\object{Q 1151$+$0651} & 18.1 & 2.758 & 900 & 0.94 & 1.17 & TN, CL & Jan. 25, 2003 & 2\\\\ \\object{Q 1209$+$0919} & 18.5 & 3.291 & 1500 & 0.78 & 1.80 & CL, TN, TK & Jan. 1, 2003 & 2\\\\ \\object{Q 1223$+$1753} & 18.1 & 2.945 & 900 & 0.66 & 1.35 & TN, CL & Jan. 25, 2003 & 2\\\\ \\object{Q 2139$-$4434} & 17.7 & 3.214 & 800 & 0.85 & 1.33 & TN & Apr. 30, 2003 & 3\\\\ \\noalign{\\smallskip}\\hline\\noalign{\\smallskip} \\multicolumn{8}{l}{{}$^\\mathrm{a}$ Legend: PH-Photometric, CL-Clear, TN-Thin cirrus, TK-Thick cirrus, WI-Windy.} \\end{tabular} \\begin{list}{}{} \\item[1:]\\citet{worseck07}: PH conditions. \\item[2:]SDSS. \\item[3:]\\citet{veron06}. \\item[4:]Slit loss corrected only. \\end{list} \\end{table*} ", "conclusions": "We have presented new evidence of the line-of-sight proximity effect as a universal phenomenon occurring in the spectra of high-redshift quasars. Even though our spectra are limited in spectral resolution, their high S/N and the power of the flux transmission method has enabled us to demonstrate the presence of the effect for every single line of sight, for the first time. Our estimate of the mean UV background intensity for the redshift range $2.71$, in agreement with theoretical expectation in $\\Lambda$CDM model. We combine the ISW likelihood function with weak lensing of CMB (hereafter Paper II \\cite{paperII}) and CMB power spectrum to constrain the equation of state of dark energy and the curvature of the Universe. While ISW does not significantly improve the constraints in the simplest 6-parameter flat $\\Lambda$CDM model, it improves constraints on 7-parameter models with curvature by a factor of 3.2 (relative to WMAP alone) to $\\Omega_K=-0.004^{+0.014}_{-0.020}$, and with dark energy equation of state by 15\\% to $w=-1.01^{+0.30}_{-0.40}$ [posterior median with ``$1\\sigma$'' (16th--84th percentile) range]. A software package for calculating the ISW likelihood function can be downloaded at {\\tt http://www.astro.princeton.edu/\\~{}shirley/ISW\\_WL.html}. ", "introduction": "Introduction} The Cosmic Microwave Background (CMB) has provided us with a wealth of cosmological information. The large-scale anisotropies were first discovered by the Differential Microwave Radiometer (DMR) on Cosmic Background Explorer (COBE) satellite \\cite{smoot92}, and the smaller-scale CMB anisotropies were subsequently measured by various ground-based/balloon-borne experiments. More recently, the Wilkinson Microwave Anisotropy Probe (WMAP) satellite \\cite{bennett03, jarosik07} produced a cosmic variance limited map of CMB anisotropies down to $l\\sim 400$. The structure of the angular power spectrum when combined with other cosmological probes (such as Sloan Digital Sky Survey, \\cite{tegmark06}, Hubble Key Project \\cite{freedman94} and 2dF Galaxy Redshift Survey \\cite{cole05}), allows extremely precise measurements of the cosmological parameters of the $\\Lambda$CDM model. While most of the fluctuations seen by WMAP and other CMB experiments were generated at the last surface of scattering, structures formed at low redshift also leave imprints on the CMB. These anisotropies, such as the thermal Sunyaev-Zeldovich (tSZ) \\cite{sunyaev80a} and kinetic Sunyaev Zeldovich effects (kSZ) \\cite{sunyaev80b}, the Integrated Sachs-Wolfe (ISW) effect \\cite{sachs67}, and gravitational lensing, contribute only slightly to the CMB power spectrum on scales measured by WMAP, but they can be detected by cross-correlating the CMB with suitable tracers of the large scale structure. This is the first of two papers that measure the Integrated Sachs-Wolfe effect and gravitational lensing (Paper II) in cross-correlation. In this paper, we focus on large scale galaxy-temperature correlations and their large scale cosmological source, the Integrated Sachs-Wolfe (ISW) effect. The ISW effect results from the red- or blue-shifting of the CMB photons as they propagate through gravitational potential wells. As the potential wells of the Universe (i.e., the spatial metric) evolve, the energy gained by photons falling into the potential well does not cancel out the energy loss as photons climb out of the well. This is important at late times when the Universe is not matter dominated and the gravitational potential is time dependent. It is only significant on large scales, since on small scales the amount of time spent by the photon in each coherence region of the gravitational potential is small and any small scale fluctuations will be smoothed out as the photon go through numerous potential wells along the line of sight. To measure the above effect, we cross-correlate the CMB temperature anisotropies with maps of galaxies from the Two Micron All Sky Survey (2MASS), luminous red galaxies (LRGs) and quasars from the Sloan Digital Sky Survey, and radio sources from the NRAO VLA Sky Survey (NVSS). This incorporates most of the LSS tracers used by previous efforts \\cite{boughn98,fosalba03,scranton03,afshordi04,boughn04,fosalba04,nolta04,padmanabhan05ISW,gaztanaga05,cabre06,giannantonio06,vielva06, pietrobon06,mcewen07,rassat07} to detect the ISW effect. Our goal in this work extends this previous literature by going beyond detecting the ISW effect to measuring its redshift evolution and using that to constrain different cosmological models (e.g. the ISW effect due to spatial curvature occurs at significantly higher redshifts than that due to a cosmological constant). We therefore require a large redshift range ($z \\sim 0$ to $2.5$) but with sufficient redshift resolution to unambiguously discern any redshift evolution of the signal. In addition, to draw robust cosmological conclusions from an observed redshift evolution, we must constrain both the redshift distribution and evolution of the bias with redshift for each of the samples; the simple assumption of constant bias is in most cases no longer sufficient. These considerations drive our survey selections; we discuss these in more detail in Sec.~\\ref{sec:discussion}. Our final product is a likelihood code that can be applied to any cosmological model. In addition to providing complementary constraints on standard cosmological parameters, we expect it can be a strong discriminator of the modified gravity models, which have very distinctive ISW predictions \\cite{song07}. We review the theory behind the ISW effect in Sec.~\\ref{sec:theory}. The CMB and LSS data sets used are described in Sec.~\\ref{sec:data}; the results of cross-correlating the two are in Sec.~\\ref{sec:cross}. Sec.~\\ref{sec:dndz} and ~\\ref{sec:sys} constrain the redshift distributions of the samples, and possible systematic contamination of the cross-correlations. Sec.~\\ref{sec:cosmo} presents the cosmological implications of these results, and Sec.~\\ref{sec:discussion} summarizes our conclusions. The companion paper (Paper II) uses the same data sets to detect the weak lensing of the CMB. All of the theoretical predictions are made with WMAP 3 year parameters ($\\Omega_b h^2$=$0.0223$, $\\Omega_c h^2$ = $0.128$, $\\Omega_K=0$, $h=0.732$, $\\sigma_8 =0.761$) except in Section~\\ref{sec:dndz} or otherwise stated. ", "conclusions": "\\label{sec:discussion} The main goal of this paper is to perform a full analysis of the integrated Sachs-Wolfe effect using the cross-correlations between WMAP CMB maps and maps of large scale structure. In contrast to previous work on this subject we place less emphasis on establishing a detection of ISW and more emphasis on developing a tool with which cosmological models can be compared to the data in a close to optimal fashion. For this reason we only select the data sets that can be reliably used towards this goal, as discussed in more detail below. The redshift range of the datasets we use is between 0 and 2.5. We use optimal weighting of the data both in angular space and in redshift space to extract the maximum amount of information, taking into account properly the correlations between them. Our final product is the likelihood function with which different cosmological models can be compared to each other. As the ISW effect is both a probe of cosmological parameters and a consistency test of the standard $\\Lambda$CDM cosmology, there have been significant previous efforts made to detect it. A number of different analysis methods have been used and the WMAP data have been cross-correlated with several samples. These include the 2MASS XSC; several SDSS samples including magnitude-limited galaxy samples, LRGs, and quasars; the NVSS; and the HEAO hard X-ray map. Most of these samples (or samples with similar spatial coverage and redshift range) are included in the present work, but not all. Here we compare our analysis with the previous work and comment on the reasons for our choice of data sets. \\newcounter{oisw} \\begin{list}{\\arabic{oisw}. }{\\usecounter{oisw}} \\item {\\em Near-infrared galaxies (2MASS).} The 2MASS galaxies are useful for ISW due to high sky coverage and the ability to see closer to the Galactic plane in the near-IR than in the optical. However they can only probe the lowest redshifts ($z<0.2$). Afshordi et~al. \\cite{afshordi04} and Rassat et~al. \\cite{rassat07} have measured the ISW signal using the 2MASS sample and we deliberately cut our 2MASS sample into brightness bins such as theirs so that we can compare the results. We find that our measured signal from 2MASS is very similar. We do however derive cosmological constraints using a Markov chain (which fits all the cosmological parameters instead of just $\\Omega_\\Lambda$) from these samples. We also take into account (albeit in a crude way) the redshift dependence of the bias resulting from seeing all nearby galaxies but only the brightest and most biased galaxies at $z\\ge 0.1$. \\item {\\em Optical galaxies (SDSS, APM).} Wide-angle multicolor galaxy surveys such as SDSS open almost limitless possibilities for constructing galaxy samples, and many of these samples have been used in previous ISW work. Most work so far has been on either flux-limited samples \\cite{fosalba03, cabre06}, which have a broad redshift distribution, or photometric LRGs \\cite{fosalba03, scranton03, padmanabhan05ISW, cabre06}, which can be seen to larger distances and for which it is easier to construct reliable photo-$z$ cuts. In SDSS, photometric LRG samples oversample the linear density field in the redshift range $0.21$), and we therefore do not confirm that the 2.1$\\sigma$ signal seen in \\cite{giannantonio06} comes from $z>1$. \\item {\\em Radio sources (NVSS).} There have been several past WMAP$\\times$NVSS ISW analyses \\cite{boughn04, nolta04, pietrobon06, mcewen07}, taking advantage of the high redshift (compared to most optical samples) and wide sky coverage of the NVSS. We have used the angular power spectrum whereas the previous works have used correlation functions or wavelet coefficients. However, the most important difference between our analysis and the previous result is that we fit $b*dN/dz$ from cross-correlations rather than using the Dunlop \\& Peacock model \\cite{dunlop90} for the redshift distribution and assuming constant bias. This is important as we find the fit $b*dN/dz$ looks very different (see Fig.~\\ref{fig:fnvss}). All of these studies, including ours, have found positive cross correlations at the $\\sim 3\\sigma$ level. However, the interpretation of this result depends sensitively on one's ability to measure $b*dN/dz$ and this is where we believe our analysis is an improvement upon previous efforts. \\item {\\em Hard X-ray background (HEAO).} Boughn \\& Crittenden \\cite{boughn04} have used the HEAO hard X-ray map \\cite{boldt87} for ISW cross-correlation. The background is due mainly to unresolved (by HEAO) active galactic nuclei and hence traces large scale structure at redshifts of order unity. This, combined with the all-sky nature of HEAO, is beneficial for ISW projects. However, we decided not to add in HEAO sample to our analysis for several reasons. First is the difficulty in understanding the $b(z)*dN/dz$ of the sample (we use the general notation $dN/dz$ here even though for unresolved X-ray flux it would be more accurate to write $dF/dz$). Only $\\sim 75$\\% of the background is resolved by Chandra into sources with measured redshifts \\cite{cowie03, boughn04b}, and we have little guidance on where to place the other 25\\%. Even if we knew $dN/dz$ perfectly, this does not tell us $b*dN/dz$: the modeled $dN/dz$ spans the range $030~\\%}$ as a prerequisite for the formation of massive strongly bound and long-term stable SCs, i.e. for the formation of young GCs \\citep{Brown+95,Burkert+96,ElmegreenEfremov97,Li+04}. On a global scale, SF efficiencies in normal spiral and irregular galaxies, as well as in starbursting dwarf galaxies are of order $0.1 - 3 \\% $ \\citep{Krueger+95}. On the smaller scale of molecular clouds in the Milky Way, the SF efficiency is of the same order of magnitude and so is the mass ratio between the molecular cloud core and the entire molecular cloud. No GC formation is therefore expected in spirals, irregulars or star-bursting dwarf galaxies by today. In giant gas-rich interacting galaxies, on the other hand, SF efficiencies of order $10 - 50 \\%$ are reported on global scales, and of order $30 - 90 \\%$ on nuclear scales of a few hundred pc up to $\\sim 1$ kpc. In those systems, GC formation should be possible. The fact that different submm lines (CO(1-0), HCN(1-0), CS(1-0)) trace molecular gas at different densities (${\\rm n \\geq 100, n~\\geq 3 \\cdot 10^4, n~\\sim 10^5~cm^{-3}}$), has allowed to see that while in the Milky Way and other nearby galaxies only a small fraction ($0.1 - 3 \\%$) of the mass of a molecular cloud makes up its high density core, the situation is drastically different in Ultraluminous Infrared galaxies (ULIRGs), which all are late stages of massive gas-rich mergers with strong nuclear (few 100 pc) starbursts. In those ULIRGs, almost all the molecular gas in the central starburst region is at the high densities of molecular cloud {\\em cores}, indicating that the molecular cloud structure must be very different from what we know in our Galaxy. The entire nuclear region is just one supergiant molecular cloud core, seriously raising the question whether the star and SC formation processes can be the same as in normal galaxies, not to mention the situation in extended, expanding, low-density tidal structures in the outskirts of other interacting galaxies. In any case, before ALMA becomes operational, the YSCs forming in these different types of environments are our best proxy to the molecular cloud structure. In the Milky Way, molecular cloud cores, molecular clouds, and YSCs all feature power-law mass functions, suggesting scale-free self-similar evolution. Not even for the closest massive merger, the Antennae, can we presently determine the molecular cloud or cloud core mass functions (cf. \\cite*{Wilson+03}). The masses of YSCs and the shape of their mass function is all we can access \\citep{Wilson+06,Anders+07}. \\cite*{GaoSolomon04} and \\cite*{Solomon+92} have shown that for all galaxies -- from Blue Compact Dwarfs to spirals and ULIRGs -- there is a tight correlation between SFR, as derived from far-infrared luminosity, and the mass in molecular cloud cores, as derived from the HCN luminosity. They also find the SF efficiency to be proportional to the mass ratio of molecular gas at core and normal densities, i.e. to the ratio between HCN or CS luminosity and CO luminosity. The highest density molecular gas in all these environments is transformed into stars with almost 100 \\% efficiency and it is the amount of gas at those high densities that controls SF. The fraction of molecular gas at the highest densities therefore defines the SF efficiency. The high ambient pressure building up in the course of massive gas-rich mergers can drive up SF efficiencies by $1-2$ orders of magnitude by compressing molecular clouds, increasing their masses and, in particular, their core mass fractions. \\cite*{JogDas92,JogDas96} have shown that the ISM pressure during mergers can easily become $3-4$ times higher than the typical internal molecular cloud pressure, raising the SF efficiency to $70-90 \\%$. This leads us to expect that the relative amount of SF that goes into the formation of massive, strongly bound young GCs in relation to the amount of SF that goes into field stars and low-mass, short-lived clusters is enhanced in massive gas-rich mergers. ", "conclusions": "{\\bf We so far conclude that starbursts in non-interacting dwarf galaxies do not form substantial populations of massive, long-lived YSCs that could evolve into GC, while massive gas-rich mergers do.} NGC 7252 is not the only example of a merger that no doubt has produced a new generation of GCs. The $\\sim 1 - 3$ Gyr old merger remnants NGC 3921 \\citep{Schweizer+96}, NGC 34 \\citep{SchweizerSeitzer07}, and NGC 1316 \\citep{Goudfrooij+01a,Goudfrooij+01b,Goudfrooij+04,Goudfrooij+07} as well feature young GC populations formed during the mergers. The metallicities of these newly formed GCs agree well with expectations on the basis of spiral galaxy ISM properties. Evolutionary synthesis models predict these SCs to take on the optical colours of the red-peak GC widely observed in E/S0 galaxies by the time the tidal features indicative of the merger origin will have vanished. They also predict that they should readily be detectable against other populations of red GCs in combined optical and NIR observations \\citep{Fritze04} (see also R. Kotulla, this volume). No example of a clearly merging gas-rich dwarf galaxy pair, nor of an accretion of a gas-rich dwarf by an elliptical or S0 has as yet been studied to check whether those would also give rise to new GC populations." }, "0801/0801.0853_arXiv.txt": { "abstract": "We discuss a cosmology in which cold dark matter begins to decay into relativistic particles at a recent epoch ($z < 1$). We show that the large entropy production and associated bulk viscosity from such decays leads to an accelerating cosmology as required by observations. We investigate the effects of decaying cold dark matter in a $\\Lambda = 0$, flat, initially matter dominated cosmology. We show that this model satisfies the cosmological constraint from the redshift-distance relation for type Ia supernovae. The age in such models is also consistent with the constraints from the oldest stars and globular clusters. Possible candidates for this late decaying dark matter are suggested along with additional observational tests of this cosmological paradigm. ", "introduction": "Understanding the nature and origin of both the dark energy \\cite{garnavich} and cold dark matter \\cite{Feng06} constitutes a significant challenge to modern cosmology. The simplest particle physics explanation for the cold dark matter is, perhaps, that of the lightest supersymmetric particle, an axion, or a heavy (e.g. \"sterile\") neutrino. The dark energy, on the other hand is generally attributed to a cosmological constant, or possibly a vacuum energy in the form of a \"quintessence\" scalar field \\cite{wetterich,ratra} or $k$-essence \\cite{zlatev,steinhardt99} which must be slowly evolving along an effective potential. See \\cite{Barger06} for a review. In addition to these explanations, however, the simple coincidence that both of these unknown entities currently contribute comparable mass energy toward the closure of the universe begs the question as to whether they could be different manifestations of the same physical phenomenon. Indeed, suggestions along this line have been made by many \\cite{Fabris,Chaplygin,Abramo04,Umezu}. See \\cite{Bean05} for a recent review. In Ref.~\\cite{Wilson07} yet another mechanism was considered by which a dark-matter particle could produce the cosmic acceleration. In that work it was shown that entropy production and an associated bulk viscosity could result from a decaying dark-matter particle. Moreover, that bulk viscosity would act as a negative pressure similar to a cosmological constant or quintessence. However, in that paper it was shown the decay alone was not sufficient to produce the observed cosmic acceleration. It was proposed, however, \\cite{Wilson07} that some, but not all of the desired cosmic acceleration could be accounted for if the particle decay was delayed by proceeding thorough a cascade of long lived intermediate states before the final entropy-producing decay. In this paper we expand on the hypothesis that the dark energy could be produced from a delayed decaying dark-matter particle. Here, we show that a dark-matter particle which, though initially stable, begins to decay to relativistic particles near the time of the present epoch will produce a cosmology consistent with the observed cosmic acceleration deduced from the type Ia supernova distance-redshift relation without the need for a cosmological constant. Hence, this paradigm has the possibility to account for the apparent dark energy without the well known fine tuning and smallness problems \\cite{Bean05} associated with a cosmological constant. Also, for the model proposed herein, the apparent acceleration is a temporary phenomenon. This avoids the difficulties \\cite{hellerman} in accommodating a cosmological constant in string theory. The idea of delayed dark matter decay is not new. It was previously introduced \\cite{Turner} as a means to provide an $\\Omega_M = 0.1-0.3$ without curvature ( $\\Omega_{tot} = 1$) by hiding matter in weakly interacting relativistic particles. Here, we point out that such a cosmology not only allows for a flat cosmology with low apparent cold dark matter matter content, but can produce an accelerating cosmology consistent with observations. We show that the bulk viscosity produced during the decay will briefly accelerate the cosmic expansion as matter is being converted from nonrelativistic to relativistic particles. This model thus shifts the dilemma in modern cosmology from that of explaining dark energy to one of explaining how an otherwise stable heavy particle might begin to decay at a late epoch. In the next section we summarize the cosmology of late decaying dark matter and its associated bulk viscosity. In the following section we discuss candidate particles for such decays, and in Section IV we present fits to the supernova magnitude-redshift relation which show that these data can be reasonably well fit in a flat $k=0$, $\\Lambda = 0$ cosmology with recent dark-matter decay. We summarize the constraints that the supernova data places on the properties of the decay along with independent constraint from the ages of oldest stars and globular clusters. ", "conclusions": "We have considered models in which the present cosmic acceleration derives from the temporary insertion of dissipative mass energy due to the bulk viscosity created by the recent decay of a cold dark-matter particle into light (undetectable) relativistic species. As illustrative examples we have considered initially matter dominated flat ($\\Lambda=0$) cosmologies plus late-time particle decay. We find that models with bulk viscosity from late-time dark-matter decay are consistent with observations of the supernova magnitude-redshift relation, and ages from the oldest stars and globular clusters. We argue that it will likely satisfy other constraints as well. Moreover, there is a difference in the SNIa magnitude-redshift relation for this cosmology compared to the standard $\\Lambda$CDM model. This is because the deceleration is faster at high redshift due to a higher matter content during the matter-dominated epoch. Thus, as more data are accumulated at the highest redshifts, it may be possible to distinguish between this cosmology and a standard $\\Lambda$CDM model. For now, however, there is sufficient success in the present model to motivate further work. In a subsequent paper we plan to consider higher order terms in the transport equation in detail as well as the effects of such late decays on the observed power spectrum of the cosmic microwave background \\cite{WMAP}, and the growth of large scale structure. Ultimately, of course, one must decide whether the dilemma of a cosmological constant with all of its difficulties is less palatable than the dilemma of a bulk viscosity produced by the delayed onset of dark-matter decay. For now, however, our purpose has simply been to establish that such a possibility exists and that it warrants further investigation." }, "0801/0801.2584_arXiv.txt": { "abstract": "With the Navy Prototype Optical Interferometer (NPOI), the binary system $\\theta^{1}$ Orionis C, the most massive member of the Trapezium, was spatially resolved over a time period extending from February 2006 to March 2007. The data show significant orbital motion over the 14 months, and, after combining the NPOI data with previous measurements of the system from the literature, the observations span 10 years of the orbit. Our results indicate that the secondary did not experience an unusually close periastron passage this year, in contradiction to the prediction of a recently published, highly eccentric $\\sim$11 year orbit. Future observations of this source will be required to improve the orbital solution. Possible implications of the results in terms of system distance are discussed, although a main conclusion of this work is that a definitive orbit solution will require more time to obtain sufficient phase coverage, and that the interaction effects expected at periastron did not occur in 2007. ", "introduction": "Orion is the nearest example of a giant molecular cloud and is the site of both high-mass star formation and a prodigious number of recently-formed stars; the central 2.5 pc region of the Orion Nebula Cluster (ONC) alone contains $\\sim$3500 stars with a combined mass exceeding 900 $M_{\\odot}$ \\citep{hil97}. Although star-forming regions such as Taurus are closer, these regions are dark clouds associated with low-mass star formation and far fewer total stars. Initial star-count studies \\citep{lad91} implied that up to 95\\% of stars formed in clustered environments like Orion, while more recent {\\it Spitzer Space Telescope} results indicate that at least half of all stars originate in dense regions \\citep{meg04}. An analysis of binary star distributions \\citep{pat02} suggests that approximately 70\\% of field stars may have formed in a clustered environment. Thus, the study of Orion, the closest giant star forming cluster, is central to investigating the early history of the majority of stars. Furthermore, the variety of cluster morphologies investigated with {\\it Spitzer} observations suggest that OB stars play a significant role in the formation and evolution of star formation in clusters \\citep{meg04}. The distance to Orion is a critical parameter that influences the interpretation of many of the properties of the region and its members. Measurements range from $\\sim$390 pc to $\\sim$480 pc using a variety of methods and assumptions including observations of water masers, radio sources, an eclipsing binary, and statistical analysis \\citep[and references therein]{gen81, sta04, san07, men07, jef07}. A closer distance would imply that the stars are less luminous, and members that are still contracting onto the main sequence are consequently older based on comparison with theoretical evolutionary tracks \\citep[e.g.,][]{sie00, pal99}. Older ages for the stars above the main sequence suggest a spread in ages. Binary stars yield a model-independent distance with the combination of a spatially resolved orbit and a double-lined radial velocity orbit \\citep[e.g.,][]{tor97}. Given the distance to Orion, very high angular resolution is required to separate binary systems which exhibit significant orbital motion. The binary system $\\theta^{1}$ Ori C (HD 37022) in the Trapezium region of Orion has such a close separation that it was not detected until speckle observations barely resolved the pair \\citep{wei99} with a separation of less than the diffraction limit of the 6m telescope employed in the discovery. The separation has decreased over time and is now best monitored with interferometry; a recent orbit fit suggested the system might be just past periastron and completing an orbital cycle within a year \\citep{kra07}. In this letter, we present the results of interferometric observations with NPOI that add significantly to the binary orbit phase coverage and suggest the orbital period may be substantially longer than predicted. The distance to $\\theta^{1}$ Ori C in particular and its physical parameters such as mass and age are of great importance since the O7 primary \\citep[e.g.,][]{gar82}, as the most massive member of the ONC, has the dominant role in shaping the properties of the surrounding nebula, and strongly impacts the circumstellar material around the ONC stars. The photoionizing radiation from $\\theta^{1}$ Ori C produces the brightening of many proplyds \\citep{ode93}, but also causes the material to escape. Observations of mass loss rates \\citep{joh98, hen99} and theoretical modeling of the results \\citep[e.g.,][]{sto99} imply that these structures cannot survive for $\\ga10^5$ yrs, substantially less than the $<$1$-$2 Myr age of the ONC \\citep{hil97}. Possible explanations for the apparent contradiction in disk lifetimes and stellar ages include radial orbits for the proplyds \\citep{sto99} or a very young age for $\\theta^{1}$ Ori C, such as has been proposed for $\\theta^{1}$ Ori B \\citep{pal01}. In contrast, recent models of the evolution of disk sizes in the ONC \\citep{cla07} suggest that disk survival times of 1-2 Myr in the uv field produced by $\\theta^{1}$ Ori C are possible -- consistent with the age of the stellar population. Key to estimating the age of $\\theta^{1}$ Ori C is placing the secondary accurately on the H-R diagram with a well-measured distance, luminosity, and mass given that the primary has already contracted onto the main sequence. We present new magnitude differences which will aid in the assessment of the luminosity, but concentrate on the orbital motion which is required to estimate the distance. ", "conclusions": "" }, "0801/0801.0638_arXiv.txt": { "abstract": "Recent astronomical observations indicate that the Universe is in the phase of accelerated expansion. There are many cosmological models which explain this phenomenon, but should we prefer those models over the simplest one -- $\\Lambda$CDM model? According to the Occam's razor principle if all models describe the observations equally well we should prefer the simplest one. We consider the model comparison methods which involve such rules: the Akaike information criterion (AIC), Bayesian information criterion (BIC) and Bayesian evidence to compare the $\\Lambda$CDM model with its generalisation where the interaction between dark matter and dark energy is allowed. The analyses based on the AIC and Bayesian evidence indicate that there is only a weak evidence in favour of the $\\Lambda$CDM model over its generalisation, while those based on BIC quantity indicate the strong evidence in favour the simpler model. We also calculate some quantity which measure the effective number of model parameters that the given data can constrain. This value is used to compare the concordance LCDM model with its generalization basing on the extended interpretation of continuity condition-interacting $\\Lambda$CDM cosmology. We conclude that data set are not enough informative to constrain all parameters allows to vary in the models. ", "introduction": "Recent observations of type Ia supernova (SNIa) provide the main evidence that current Universe is in an accelerating phase of expansion \\cite{Riess:1998cb}. Cosmic microwave background (CMB) data indicate that the present Universe has also negligible space curvature \\cite{Spergel:2006hy}. Therefore if we assume the Friedmann-Robertson-Walker (FRW) model in which effects of nonhomogeneities are neglected, then the acceleration must be driven by dark energy component $X$ (matter fluid violating the strong energy condition $\\rho_{\\text{X}}+3p_{\\text{X}}\\geq 0)$. This kind of energy represents roughly $70\\%$ of the matter content of the current Universe. Because the nature as well as mechanism of the cosmological origin of the dark energy component are unknown some alternative theories try to eliminate the dark energy by modifying the theory of gravity itself. The main prototype of this kind of model is covariant brane models based on the Dvali-Gabadadze-Porrati (DGP) model \\cite{Dvali:2000hr} as generalized to cosmology by Deffayet \\cite{Deffayet:2000uy}. The simplest explanation of a dark energy component is the cosmological constant with effective equation of state $p=-\\rho$ but appears the problem of its smallness and hence its relatively recent dominance. Although the $\\Lambda$CDM model offers possibility of explanation of observational data it is only effective theory which contain the enigmatic theoretical term -- the cosmological constant $\\Lambda$. Other numerous candidates for dark energy description have also been proposed like to evolving scalar field \\cite{Peebles:1987ek} usually referred as quintessence, the phantom energy \\cite{Caldwell:1999ew, Dabrowski:2003jm}, the Chaplygin gas \\cite{Kamenshchik:2001cp} etc. Some authors believed that the dark energy problem belongs to the quantum gravity domain \\cite{Witten:2000zk}. These theoretical models are consistent with the observations, they are able to explain the phenomenon of the accelerated expansion of the Universe. But should we really prefer such models over the $\\Lambda$CDM one? To answer this question we should use some model comparison methods to confront existing cosmological models having observations at hand. Let us assume that we have $N$ pairs of measurements $(y_i,x_i)$ and that we want to find the relation between the $y$ and $x$ quantities. Suppose that we can postulate $k$ possible relations $y\\equiv f_i(x,\\bar{\\theta})$, where $\\bar{\\theta}$ is the vector of unknown model parameters and $i=1,\\dots,k$. With the assumption that our observations come with uncorrelated gaussian errors with mean $\\mu_i=0$ and standard deviation $\\sigma_i$ the goodness of fit for the theoretical model is measured by the $\\chi^2$ quantity given by \\begin{equation} \\chi^2=\\sum_{i=1}^{N} \\frac{(f_l(x_i,\\bar{\\theta}) - y_i)^2}{2\\sigma_i^2}=-2\\ln L, \\end{equation} where $L$ is the likelihood function. For the particular family of models $f_l$ the best one minimize the $\\chi^2$ quantity, which we denote $f_l(x,\\hat{\\bar{\\theta}})$. The best model from our set of $k$ models ${f_1(x,\\hat{\\bar{\\theta}}),\\dots,f_k(x,\\hat{\\bar{\\theta}})}$ could be the one with the smallest value of $\\chi^2$ quantity. But this method could give us misleading results. Generally speaking for more complex model the value of $\\chi^2$ is smaller, thus the most complex one will be choose as the best from our set under consideration. A clue is given by so called Occam's razor principle: ``If two models describe the observations equally well, choose the simplest one.'' This principle has aesthetic as well as empirical justification. Let us quote simple example which illustrate this rule \\cite{MacKay:2003} and present it in Figure \\ref{Fig:1}. We see the black box and the white one behind it. One can postulate two models: 1. There is one box behind the black box, 2. There are two boxes of identical height and colour behind the black box. Both models explain our observations equally well. According to the Occam's razor principle we should accept the explanation which is simpler so that there is only one white box behind the black one. Is not it more probable that there is only one box than two boxes with the same height and colour? \\begin{figure}[h] \\centering \\includegraphics[scale=0.5]{fig1.eps} \\caption{Illustration to the example explaining the Occam's razor principle.} \\label{Fig:1} \\end{figure} We could not use this principle directly because the situations when two models explain the observations equally well are rare. But in the information theory as well as in the bayesian theory there are methods for model comparison which include such rule. In the information theory there are no true models. There is only reality which can be approximated by models, which depend on some number of parameters. The best one from the set under consideration should be the best approximation to the truth. The information lost when truth is approximated by model under consideration is measured by so called Kullback-Leibler (KL) information so the best one should minimizes this quantity. It is impossible to compute the KL information directly because it depends on truth which is unknown. Akaike \\cite{Akaike:1974} found approximation to the KL quantity which is called the Akaike information criterion (AIC) and is given by \\begin{equation} \\text{AIC}=-2\\ln \\mathcal{L} +2d, \\end{equation} where $\\mathcal{L}$ is the maximum of the likelihood function and $d$ is the number of model parameters. Model which is the best approximation to the truth from the set under consideration has the smallest value of the AIC quantity. It is convenient to evaluate the differences between the AIC quantities computed for the rest of models from our set and the AIC for the best one. Those differences ($\\Delta_{\\text{AIC}}$) are easy to interpret and allow a quick `strength of evidence' for considered model with respect to the best one. The models with $0 \\le \\Delta_{\\text{AIC}}\\le 2$ have substantial support (evidence), those where $4<\\Delta_{\\text{AIC}}\\le 7$ have considerably less support, while models having $\\Delta_{\\text{AIC}} > 10 $ have essentially no support with respect to the best model. It is worth noting that the complexity of the model is interpreted here as the number of its free parameters that can be adjusted to fit the model to the observations. If models under consideration fit the data equally well according to the Akaike rule the best one is with the smallest number of model parameters (the simplest one in such an approach). In the Bayesian framework the best model (from the model set under consideration) is that which has the largest value of probability in the light of data (so called posterior probability) \\cite{Jeffreys:1961} \\begin{equation} P(M_{i}|D)=\\frac{P(D|M_{i})P(M_{i})}{P(D)}, \\end{equation} where $P(M_{i})$ is a prior probability for the model $M_{i}$, $P(D)$ is normalization constant [$P(D)= \\sum _{i=1}^{k} P(D|M_{i})P(M_{i})$], $D$ denotes data. $P(D|M_{i})$ is the marginal likelihood, also called evidence \\begin{equation} P(D|M_{i})=\\int P(D|\\bar{\\theta},M_{i})P(\\bar{\\theta}|M_{i}) d \\bar{\\theta} \\equiv E_{i}, \\end{equation} where $P(D|\\bar{\\theta},M_{i})$ is likelihood under model $i$, $P(\\bar{\\theta}|M_{i})$ is prior probability for ${\\bar{\\theta}}$ under model $i$. Let us note that we can include the Occam's razor principle by assuming the greater prior probability for simpler model, but this is not necessary and rarely used in practice. Usually one assume that there is no evidence to favor one model over another which cause to equal value of prior for all models under consideration. It is convenient to evaluate the posterior ratio for models under consideration which in the case with flat prior for models is reduced to the evidence ratio (so called the Bayes factor $B$). The interpretation of the logarithm of Bayes factor values are as follows: $0<\\ln B\\leq 2$ as a weak, $2<\\ln B\\leq 6$ as a positive, $6<\\ln B\\leq 10$ as a strong and $\\ln B> 10$ as a very strong evidence in favor better model. This quantity involves the Occam's razor rule. Let us simplify the problem to illustrate how this principle works here \\cite{MacKay:2003,Trotta:2005ar}. Assume that $\\bar{P}(\\bar{\\theta}|D,M)$ is the non normalized posterior probability for the vector $\\bar{\\theta}$ of model parameters. In this notation $E=\\int\\bar{P}(\\bar{\\theta}|D,M)d\\bar{\\theta}$. Suppose that posterior has a strong pick in the maximum: $\\bar{\\theta}_{\\text{MOD}}$. It is reasonable to approximate the logarithm of the posterior by its taylor expansion in the neighbour of $\\bar{\\theta}_{\\text{MOD}}$ so we finished with the expression \\begin{equation} \\bar{P}(\\bar{\\theta}|D,M)=\\bar{P}(\\bar{\\theta}_\\text{MOD}|D,M)\\exp \\left[-(\\bar{\\theta}-\\bar{\\theta}_{\\text{MOD}})^T\\text{C}^{-1}(\\bar{\\theta}-\\bar{\\theta}_{\\text{MOD}})\\right], \\end{equation} where $\\left[\\text{C}^{-1}\\right]_{ij}=-\\left[\\frac{\\partial^2\\ln\\bar{P}(\\bar{\\theta}|D,M)}{\\partial\\theta_i\\partial\\theta_j}\\right]_{\\bar{\\theta}=\\bar{\\theta}_\\text{MOD}}$. Posterior is approximated by the gaussian distribution with the covariance matrix $\\text{C}$ and the mean $\\bar{\\theta}_\\text{MOD}$. The expression for the evidence take a form $E=\\bar{P}(\\bar{\\theta}_{\\text{MOD}}|D,M)\\int\\exp \\left[-(\\bar{\\theta}-\\bar{\\theta}_{\\text{MOD}})^T\\text{C}^{-1}(\\bar{\\theta}-\\bar{\\theta}_{\\text{MOD}})\\right]d \\ \\bar{\\theta}.$ Because the posterior has a strong pick near the maximum, the most contribution to the integral comes from the neighbour close to $\\bar{\\theta}_\\text{MOD}$. Contribution from the other other region of $\\bar{\\theta}$ can be ignored, so we can expand the limit of the integral to whole $R^d$. With this assumptions one can obtain $E=(2\\pi)^{\\frac{d}{2}}\\sqrt{\\det\\text{C}}\\bar{P}(\\bar{\\theta}_{\\text{MOD}}|D,M)= (2\\pi)^{\\frac{d}{2}}\\sqrt{\\det\\text{C}}P(D|\\bar{\\theta}_\\text{MOD},M)P(\\bar{\\theta}_\\text{MOD}|M)$. Suppose that the likelihood function has sharp pick in $\\hat{\\bar{\\theta}}$ and the prior for $\\bar{\\theta}$ is nearly flat in the neighbour of $\\hat{\\bar{\\theta}}$. In this case $\\hat{\\bar{\\theta}}=\\bar{\\theta}_{\\text{MOD}}$ and the expression for the evidence takes the form $E=\\mathcal{L}(2\\pi)^{\\frac{d}{2}}\\sqrt{\\det\\text{C}}P(\\hat{\\bar{\\theta}}|M)$. The $(2\\pi)^{\\frac{d}{2}}\\sqrt{\\det\\text{C}}P(\\hat{\\bar{\\theta}}|M)$ quantity is called the Occam factor (OF). When we consider the case with one model parameter with flat prior $P(\\theta|M)=\\frac{1}{\\Delta\\theta}$ the Occam factor OF$=\\frac{2\\pi\\sigma}{\\Delta\\theta}$ which can be interpreted as the ratio of the volume occupied by the posterior to the volume occupied by prior in the parameter space. The more parameter space wasted by the prior the smaller value of the evidence. It is worth noting that the evidence does not penalize parameters which are unconstrained by the data \\cite{Liddle:2006kn}. As the evidence is hard to evaluation an approximation to this quantity was proposed by Schwarz \\cite{Schwarz:1987} so called Bayesian information criterion (BIC) and is given by \\begin{equation} \\text{BIC}=-2\\ln\\mathcal{L}+2d\\ln N, \\end{equation} where $N$ is the number of the data points. The best model from the set under consideration is this which minimize the BIC quantity. It is also convenient to analyse the difference between BIC quantities for the rest models from our set with the BIC for the best one. These differences could be interpreted in the following way: $0<\\Delta_{\\text{BIC}}\\leq 2$ as a weak, $2<\\Delta_{\\text{BIC}}\\leq 6$ as a positive, $6<\\Delta_{\\text{BIC}}\\leq 10$ as a strong and $\\Delta_{\\text{BIC}}> 10$ as a very strong evidence in favour of a better model. One can notice the similarity between the AIC and BIC quantities though they come from different approaches to model selection problem. The dissimilarity is seen in the so called penalty term: $ad$, which penalize more complex models (complexity is identified here as the number of free model parameters). One can evaluated the factor by which the additional parameter must improve the goodness of fit to be included in the model. This factor must be greater than $a$ so equal to $2$ in the AIC case and equal to $\\ln N$ in the BIC case. Notice that the latter depends on the number of the data points. It should be pointed out that presented model selection methods are widely used in context of cosmological model comparison \\cite{Hobson:2002de, Beltran:2005xd, Mukherjee:2005wg, Mukherjee:2005tr, Trotta:2005ar, Niarchou:2003hz, Liddle:2004nh, Saini:2003wq, John:2002gg, Parkinson:2004yx, John:2005bz, Serra:2007id, Biesiada:2007um, Liddle:2006kn, Godlowski:2005tw, Szydlowski:2006ay, Szydlowski:2005xv, Kunz:2006mc, Parkinson:2006ku, Trotta:2007hy, Liddle:2007fy, Kurek:2007tb}. We should keep in mind that conclusions based on such quantities depend on the data at hand. Let us mention again the example with the black box. Suppose that we made a few steps toward this box that we can see the difference between the height of the left and right side of the white box. Our conclusion change now. Let us quote example taking from \\cite{John:2005bz}. Assume that we want to compare the Newtonian and Einsteinian theories in the light of the data coming from laboratory experiment where general relativistic effects are negligible. In this situation Bayes factor between Newtonian and Einsteinian theories will be close to unity. Whereas comparing the general relativistic and Newtonian explanations of the deflection of a light ray that just grazes the sun's surface give the Bayes factor $\\sim 10^{10} $ in the favor of the first one (and even greater with more accurate data). Having this in mind an interesting supplement to the above considerations seems to be quantity which measure the effective number of model parameters (C$_b$) that the given data set can constrain \\cite{Kunz:2006mc,Liddle:2007fy}. It can be computed using the relation \\begin{equation} \\text{C}_b = \\overline{\\chi^2(\\bar{\\theta})} - \\chi^2(\\hat{\\bar{\\theta}}), \\end{equation} where the mean is taken over the posterior pdf for $\\bar{\\theta}$ and $\\hat{\\bar{\\theta}}$ is the mode of the posterior pdf (different choices are also possible). As have been shown this quantity correspond to the number of parameters for which the width of the posterior probability distribution is significantly narrower than the width of the prior probability distribution \\cite{Kunz:2006mc}. These parameters can be considered to have been well measured by the data given our prior assumption in the model. It helps to determine if the data is informative enough to measure the parameters under consideration. We share with George Efstathiou opinion \\cite{Efstathiou:2007gz,Chongchitnan:2007eb,Szydlowski:2004jv} that there is no sound theoretical basis for considering the dynamical dark energy, where as we are beginning to see an explanation for a small cosmological constant emerging from more fundamental theory. In our opinion the $\\Lambda$CDM model has the status of satisfactory effective theory. Estathiou argued why the cosmological constant should be given higher weight as a candidate for dark energy description than dynamical dark energy. In this argumentation Occam's razor is used to point out a more economical model explaining the observational data. On the other hand Biesiada advocated the use of Akaike information criterion which favour rather a dynamical model of dark energy (quintessence model) \\cite{Biesiada:2007um}. In our opinion quantification of Occam's razor by computing Bayesian evidence and Bayesian information criterion is more suitable as the AIC is unadequate for many statistical problems \\cite{Dowe:2007bn}. The main aim of this paper is to compare the simplest cosmological model -- the $\\Lambda$CDM model -- with its generalisation where the interaction between dark energy and dark matter sector is allowed using methods described above. ", "conclusions": "We presented the methods of model comparison coming from information as well as bayesian theory. Both of them include the Occam's razor principle which states that if two models describe the observations equally well we should choose the simpler one. According to Akaike and Schwarz rule the model complexity is interpreted in the term of number of free model parameter while according to the Bayesian evidence more complex model wast greater volume of the parameter space. Finally we present the quantity which measures the effective number of model parameters which the data used in analysis can constrain. This quantity is also called the Bayesian complexity and reduce to the number of free model parameters when the data set is highly informative. We use those methods to answer the question if we should prefer the generalisation of the simplest $\\Lambda$CDM model where the interaction between dark matter and dark energy sector is allowed. The AIC and bayesian evidence give similar results: there is only weak evidence to favor the $\\Lambda$CDM model over the interacting $\\Lambda$CDM one. This is with contrary with the conclusion from the analysis of the BIC quantity: here there is a strong evidence to favor the $\\Lambda$CDM model. The analysis of the Bayesian complexity gives that data sets considered carry not enough information to constraint all parameters allows to vary in the models considered." }, "0801/0801.0312_arXiv.txt": { "abstract": "We compare environmental effects in two analogous samples of galaxies, one from the Sloan Digital Sky Survey (SDSS) and the other from a semi-analytic model (SAM) based on the Millennium Simulation (MS), to test to what extent current SAMs of galaxy formation are reproducing environmental effects. We estimate the large-scale environment of each galaxy using a Bayesian density estimator based on distances to all ten nearest neighbors and compare broad-band photometric properties of the two samples as a function of environment. The feedbacks implemented in the semi-analytic model produce a qualitatively correct galaxy population with similar environmental dependence as that seen in SDSS galaxies. In detail, however, the colors of MS galaxies exhibit an exaggerated dependence on environment: the field contains too many blue galaxies while clusters contain too many red galaxies, compared to the SDSS sample. We also find that the MS contains a population of highly clustered, relatively faint red galaxies with velocity dispersions comparable to their Hubble flow. Such high-density galaxies, if they exist, would be overlooked in any low-redshift survey since their membership to a cluster cannot be determined due to the ``Fingers of God'' effect. ", "introduction": "Since the discovery of the morphology-environment relation \\citep{Dressler_1980}, it has been known that galaxy properties are correlated with their large-scale environment: the average morphology, color and luminosity of galaxies differ depending on how crowded their neighborhood is. On the face of it, it is not clear why or how the environment of a galaxy on Mpc scales should be related to the kpc-scale processes (star formation, supernova and AGN feedback) that determine the bulk properties of a galaxy. To further confuse matters, the strong correlation between morphology, color and luminosity \\citep[][and references therein]{Strateva_2001} makes it unclear which property is ultimately driven by environment let alone which physical processes are responsible. It is not even clear to what extent the environment of a galaxy effects it through nature (different formation conditions) rather than nurture (galaxy-galaxy interactions). A critical step towards answering such questions is to compare observed trends to those present in a simulated galaxy ensemble in which one knows all the processes at work. The Sloan Digital Sky Survey \\citep[SDSS,][]{Adelman_2007} is a powerful tool for addressing questions of environmental effects. Its spectroscopic sample of galaxies is the largest such sample ever, ensuring that even relatively rare galaxy populations are well represented, and the survey's large contiguous footprint makes it easy to determine the large-scale environment for most of these galaxies. Previous researchers who have used the SDSS galaxy catalog to study environmental dependences have found three broad trends: the peaks of the bimodal color distribution of galaxies do not shift for different environments; blue and red galaxies are most common in low- and high-density environments, respectively; the luminosity of red galaxies increases with local density \\citep{Hogg_2003,Kauffmann_2004,Balogh_2004,Tanaka_2004,Blanton_2005,Zehavi_2005,Park_2007, Ball_2007}. The Millennium Simulation \\citep[MS,][]{Springel_2005} is the largest ever cosmological simulation comprising some $10^{10}$ dark matter (DM) particles with a spatial resolution of $5h_{100}^{-1}$ kpc. At $z=0$, the simulation fills a cube $500h_{100}^{-1}$ Mpc per side. The MS does not explicitly model the gas, dust and stars which make up observable galaxies, but it produces a DM halo merger tree which serves as the backbone for a number of semi-analytic models (SAMs). Unlike N-Body/SPH simulations, SAMs do not simulate the gravitational and hydrodynamic forces involved in the formation and evolution of galaxies, but they do provide a computationally inexpensive way to explore the parameter space of sub-grid processes. The trade-off is that the parameters of a SAM must be tuned using observations (e.g. matching to the observed luminosity function), making truly independent comparisons between the model and reality more challenging. Numerous groups have developed SAMs which hierarchically form some $10^{7}$ galaxies from the MS merger tree \\citep[eg:][and references therein]{Croton_2006, Bower_2006, DeLucia_2007}. Their models differ ---for example in their treatment of AGN feedback--- but all reproduce some of the empirical features of galaxy populations. The MS galaxies have a very realistic distribution of luminosities, thanks to judicious use of ``radio'' feedback. They also exhibit a bimodal color distribution as discovered in SDSS \\citep{Strateva_2001}. Finally, the power spectrum of the density fluctuations is in good agreement with the empirical data from 2dF and SDSS \\citep{Springel_2005}. Previous investigators have found that the brightest MS galaxies are red, dead ellipticals populating rich galaxy clusters \\citep{DeLucia_2006}, while the modeled galaxies in the very lowest-density environments have similar colors and star formation rates as analogous SDSS galaxies \\citep{Patiri_2006}. In this work we compare the observed galaxy populations with those produced with SAMs. Our work differs from those listed above in the following ways: we use Bayesian number density as a proxy for local environment, rather than the commonly used surface density or two-point correlation function; we use the $u-r$ color, which has more leverage than the $g-r$ color; we use SDSS Data Release 5, rather than any of the previous (smaller) releases; and last but not least, we compare observed and modeled galaxies for the full range of galaxy environments and colors. ", "conclusions": "We have compared two analogous galaxy samples, one from the SDSS DR5 spectroscopic sample, and one from the SAMs of \\cite{DeLucia_2007}, after correcting for the observational effects present in the former. The density distribution and the luminosity function of the modeled galaxies qualitatively match those for the SDSS sample, but there are $50$\\% too many blue galaxies in the former. In detail, two additional discrepancies become apparent between the galaxy samples: MS galaxies are more blue in $u-r$ than SDSS galaxies; the colors of galaxies depend more strongly on environment in MS than in SDSS. The strong environmental dependence manifests itself as an over-representation of blue galaxies overall and suggests that the feedbacks implemented by \\cite{DeLucia_2007} exaggerate the role of galaxy environment. A population of relatively faint red galaxies in extremely high density environments is visible in the MS survey \\emph{sans} observational effects. Such high density environments would be imperceptible in SDSS due to velocity dispersions comparable to local Hubble flow." }, "0801/0801.2121_arXiv.txt": { "abstract": "According to recent experimental data at GSI, the rate of the number of daughter ions ${^{140}}{\\rm Ce}^{58+}$, produced by the nuclear K--shell electron capture ($EC$) decay of the H--like ion ${^{140}}{\\rm Pr}^{58+}$, is modulated in time with a period $T_{EC} = 7.06(8)\\,{\\rm sec}$ and an amplitude $a_{EC} = 0.18(3)$. We show that this phenomenon can be explained by neutrino mass differences and derive a value for the difference of squared masses $\\Delta m^2_{21} = m^2_2 - m^2_1 = 2.22(3)\\times 10^{-4}\\,{\\rm eV}^2$. PACS: 12.15.Ff, 13.15.+g, 23.40.Bw, 26.65.+t ", "introduction": " ", "conclusions": "" }, "0801/0801.1122_arXiv.txt": { "abstract": "We present theoretical models for the formation and evolution of populations of low-mass X-ray binaries (LMXB) in the two elliptical galaxies NGC 3379 and NGC 4278. The models are calculated with the recently updated \\emph{StarTrack} code \\citep{Belczynski2006}, assuming only a primordial galactic field LMXB population. StarTrack is an advanced population synthesis code that has been tested and calibrated using detailed binary star calculations and incorporates all the important physical processes of binary evolution. The simulations are targeted to modeling and understanding the origin of the X-ray luminosity functions (XLF) of point sources in these galaxies. For the first time we explore the population XLF down to luminosities of $3\\times10^{36}\\, \\rm erg\\,s^{-1}$, as probed by the most recent observational results \\citep{Kim2006}. We consider models for the formation and evolution of LMXBs in galactic fields with different CE efficiencies, stellar wind prescriptions, magnetic braking laws and initial mass functions. We identify models that produce an XLF in excellent agreement with the observations both in shape and absolute normalization. We also find that the treatment of the outburst luminosity of transient systems remains a crucial factor for the determination of the XLF since the modeled populations are dominated by transient X-ray systems. ", "introduction": "A Low Mass X-ray binary (LMXB) is a Roche lobe overflowing, mass-transfering binary system with a compact object accretor, either a black hole (BH) or a neutron star (NS), and a low mass ($\\gtrsim 1\\,\\rm M_{\\odot}$) donor. Since the late 80's it has been suggested that LMXBs should exist in early type galaxies, E and S0, and that they might even dominate the X-ray emission \\citep{TF1985, fabbiano1989, KFT1992}. The stellar populations in these galaxies are typically old and homogeneous. Massive stars have already evolved to compact objects and LMXBs are probably the only sources with X-ray luminosities above $10^{36}\\, \\rm erg\\,s^{-1}$. Uncontroversial detection of LMXBs in early type galaxies became possible only this last decade with \\textit{Chandra's} increased angular resolution \\citep{Fabbiano2006, SIB2000}. The spectra of individual X-ray sources are consistent with those expected from LMXB models and the LMXBs observed in the Milky-way and M31 \\citep{HB2006, IAB2003}. For many galaxies observed with \\textit{Chandra}, the XLFs have been derived and they can usually be fitted with a single or a broken power law. The detections limit for these surveys is usually a few times $10^{37}\\,\\rm erg\\,s^{-1}$. \\citet{KF2004} derived XLFs for 14 early type galaxies and they included completeness corrections. Each XLF is well fitted with a single power law with cumulative slope between -0.8 and -1.2. The composite XLF of these galaxies though is not consistent with a single power law. There is a prominent break at $(5\\pm1.6)\\times 10^{38}\\, \\rm erg\\,s^{-1}$, close to the Eddington luminosity ($L_{\\rm Edd}$) of a helium accreting NS-LMXB. This break might be hidden in the individual XLFs due to poor statistics \\citep[see also][]{SIB2000, KMZ2002, Jordan2004, Gilfanov2004}. Other recent studies by \\citet{JCBG2003}, \\citet{SSI2003} and \\citet{Jordan2004} suggested a break of the XLF at a higher luminosity ($\\sim 10^{39}\\, erg\\,s^{-1}$). The exact position and the nature of these breaks are still somewhat controversial, as the correct interpretation of the observed XLFs relies significantly on the proper completeness correction when looking at luminosities close to the detection limit and small number statistics at the high-end of the XLF. Recent \\textit{Chandra} observations \\citep{Kim2006} have yielded the first low-luminosity XLFs of LMXBs for two typical old elliptical galaxies, NGC3379 and NGC4278. The detection limit in these observations is $\\sim 3\\times 10^{36}\\rm \\, erg\\, s^{-1}$ which is about an order of magnitude lower than in most previous surveys of early type galaxies. The observed XLFs of the two ellipticals extend only up to $6\\times 10^{38}\\,\\rm erg\\, s^{-1}$ and are well represented by a single power law with a slope (in a differential form) of $1.9\\pm0.1$. When \\textit{Chandra} observations are compared with optical images from \\textit{Hubble} or other ground based telescopes, it is generally found that a significant fraction of the LMXBs are inside globular clusters (GC). On average 4-5\\% of the GCs in a given galaxy are associated with a LMXB (with $L_x>\\sim 10^{37}\\rm \\,erg \\, s^{-1}$), while the fraction of LMXBs located in GC, varies from 10\\% to 70\\% depending on the type of the galaxy and its GC specific frequency. At present the origin and the properties of these systems, both in GCs and the field, are not yet well understood. It has been noted that XLFs at high luminosities for each sub-group (GCs and field) do not reveal any differences within the statistics of the samples considered \\citep{Fabbiano2006, KimE2006, KMZ2007, Jordan2004, Sarazin2003}. However, in more recent studies, \\citep{Fabbiano2007} and \\citet{VG2007} independently found that the two XLFs (GCs and field sources) show significant differences at low luminisities (below $10^{37}\\rm \\, erg\\, sec^{-1}$), pointing to a different LMXB formation mechanism in GCs. The natural question that arises is whether (\\textit{i}) all LMXBs were formed in GCs through dynamical interactions and some eventually escaped or some GCs dissolved in the field, or (\\textit{ii}) field LMXBs were born in situ through binary evolution of primordial binaries. The formation rates associated with these two possibilities are not understood well enough to give accurate predictions and use the relative numbers in the samples. \\citet{Juett2005} has shown that the observed relationship between the fraction of LMXBs found in GCs and the GC-specific frequency in early-type galaxies is consistent with the galactic field LMXB population being formed in situ. Similarly, \\citet{Irwin2005} compared the summed X-ray luminosity of the LMXBs to the number of GCs in a galaxy; in the case of all LMXBs having formed exclusively in GCs, the two should be directly proportional regardless of where the LMXBs currently reside. Instead he found that the proportionality includes an additive offset implying the existence of a LMXB population unrelated to GCs. In the past, semi-analytical theoretical models have been introduced for the study of the LMXB population in early galaxies. \\citet{White1998} studied the connection between the star formation rates of normal galaxies, i.e. galaxies without an active nucleus, and the formation rate of LMXBs and millisecond pulsars, assuming that all LMXBs are formed from primordial binaries. Considering a time-dependent star formation rate, they showed that the general relativity timescales relevant to the evolution of primordial binaries to LMXBs and to millisecond pulsars, lead to a significant time delay of the peak in the formation rate of these populations after the peak in the star formation rate. In a followup work \\citet{Ghosh2001}, using several updated star formation rate models, calculated the evolution of the X-ray luminosity of galaxies. They found that different star formation models lead to very different X-ray luminosity profiles, so the observed X-ray profiles can be used as probes of the star formation history. Finally they compared their models with the first \\textit{Chandra} deep imaging observations, to conclude that these first results were consistent with current star formation models. \\citet{Piro2002} argued that the majority of LMXBs in the field of elliptical galaxies have red giant donors feeding a thermally unstable disk and stay in this transient phase for at least 75\\% of their life. The very luminous X-ray sources ($L_x>10^{39} \\rm \\, erg\\, s^{-1}$) detected in \\textit{Chandra} surveys have been suggested to be X-ray binaries with highly super-Eddington mass inflow near the accreting component. In elliptical galaxies these objects have been suggested by King 2002 to be micro-quasar-like, as these galaxies contain no high mass X-ray binaries \\citep{King2002}. More recently \\citet{Ivanova2006} also argued that this bright end of the XLF is most likely dominated by transient LMXBs with BH accretors during outburst and it can be used to derive constraints on the BH mass function in LMXBs; they also showed that the standard assumption of a constant transient duty cycle (DC) across the whole population seems to be inconsistent with current observations. Semi-analytical population synthesis (PS) models of LMXBs have also been constructed for late type galaxies. \\citet{Wu2001} created a simple birth-death model, in which the lifetimes of the binaries are inversely proportional to their X-ray luminosity, and calculated the XLFs of spiral galaxies. His models reproduce some features, such as the luminosity break in the observed XLFs of spiral galaxies. The position of this break depends on the star formation history of the galaxy, and he suggested that it can be used as a probe of the galaxy's merger history. The formation of LMXBs in GCs via dynamical interactions is less well studied, since apart from the binary stellar evolution, one has to also take into account the complex cluster dynamics. \\citet{BD2004} considered a semi-analytical model for accretion from degenerate donors onto NSs in ultracompact binaries and showed that binaries with orbital periods of 8-10 minutes and He or C/O white dwarf (WD) donors of 0.06-0.08 $M_\\odot$ naturally provide the primary slope (-0.8 for cumulative form) typically derived from XLFs of elliptical galaxies. Ultra-compact systems are predicted to form in the dense GC environment and have relatively short persistent lifetimes ($<3\\times 10^6\\, \\rm yr$) but they form continuously through dynamical interactions. \\citet{Ivanova2007} presented PS studies of compact binaries containing NSs in dense GCs. They used \\emph{StarTrack} as their PS modeling tool in addition to a simplified treatment for the dynamical interactions. Their models produced a mixed population of LMXBs with red giant and MS donors, and ultra-compact X-ray binaries; relative formation rates can be comparable but the different sub-populations have very different lifetimes. In this paper we investigate the plausibility of an important contribution to the XLFs of these two galaxies from a primordial galactic field LMXB population using advanced PS simulations. In \\S 2 we describe briefly the physics included in our PS code and explain in detail the way we are constructing the modeled XLFs and the treatment of transient LMXBs. We discuss the results of our simulations in \\S 3: the modeled XLFs from different models, a statistical comparison with the observed XLFs of the elliptical galaxies NGC3379 and NGC4278, and an analysis of the dependence of the modeled XLF properties on the PS parameters. Finally in \\S 4 we discuss the implication of our findings and the caveats of our methods. ", "conclusions": "The recent deep \\textit{Chandra} observations \\citep{Kim2006} of the two typical old elliptical galaxies: NGC3379 and NGC4278, led to the first observed low-luminosity XLFs of LMXBs, with the detection limit ($3\\times 10^{36}\\rm \\, erg\\, s^{-1}$) being about an order of magnitude lower than in most previous surveys. Motivated by this observational work, we developed PS simulations of LMXBs appropriate for these two galaxies. We considered formation of LMXBs only through the evolution of primordial binaries in the galactic fields and examined the possible contribution to the overall LMXB population. For our modeling we used the updated PS code \\emph{StarTrack} \\citep{Belczynski2006}. Our main conclusions can be summarized as follows: We found that some of our models produce XLFs in very good agreement with the observations, based on both the XLF shape and absolute normalization. There is no unique combination of PS parameters and modeling of transient sources (DC and outburst luminosity) that gives an XLF in agreement with the observation. We conclude that formation of LMXBs in the galactic field via evolution of primordial binaries can have a significant contribution to the total population of an elliptical galaxy, especially the ones with low GC specific frequency such as NGC3379 \\citep{Fabbiano2007}. Nevertheless, we are able to exclude the majority of our models, as inconsistent with the observations. Note that widely used, simple assumptions such as that all transients source in the outburst state are emitting X-rays at $L_{\\rm Edd}$, lead to XLFs clearly inconsistent with the observed ones. Our results appear to be robust since we do not have to fine-tune our code parameters in order to get a model that resembles the observed population. As already suggested by \\citet{Piro2002}, the LMXB population has a significant contribution by transient systems (thermal disk instability) and with reasonable outburst DCs they can even dominate the XLF. As a consequence the XLF shape is rather sensitive to the treatment of these transient systems. In Figure \\ref{xlf_dc} we show that keeping the same PS parameters and changing \\emph{only} the modeling of transient sources leads to completely different XLFs. We tried different methods of modeling the outburst characteristics of transient LMXBs and we found that: \\emph{(i)} When we assume the outburst luminosity of all transient LMXBs to be equal to $L_{\\rm Edd}$ (transient treatment E) or apply eq.(\\ref{Lx_P}) (transient treatment F) - which was empirically derived for Galactic BH-LMXBs - to the whole population, we get XLFs inconsistent with the observed ones regarding both their shape and the total number of sources predicted. \\emph{(ii)} A constant DC for all systems, although not physically motivated, can be sometimes a good first approximation. \\emph{(iii)} We get the best agreement with observations, when we consider a variable DC for NS-LMXB based on the theoretical study of \\citet{DLM2006} (see eqs. \\ref{dc} and \\ref{Lx_phys}), while for the BH-LMXBs we use the empirical correlation between orbital period and outburst luminosity, derived by \\citet{PDM2005} (see eq. \\ref{Lx_P}), and assuming a low constant DC ($\\sim 5\\%$). The LMXB sub-populations that mainly contribute to the model XLFs are NS-LMXBs with red giant donors and BH-LMXBs with MS donors. A population of persistent ultra compact LMXBs with WD donors is also present in our models and in some cases has an important contribution too (see models $\\rm 9A^{IT}$, $\\rm 11A^{IT}$ and $\\rm 12A^{IT}$). Of these sub-populations, the NS-LMXBs are the most dominant and primarily determine the XLF shape in the medium and low luminosity range (below $2 \\times 10^{38}\\rm \\, erg \\, s^{-1}$), while the BH-LMXBs have a significant contribution to the high-end of the XLF. The normalization of the modeled XLFs is a less robust characteristic than its shape. We normalize the models so that the number of the primordial binaries we evolve, correspond to the known galaxy masses, given the IMF and the binary fraction. There are however uncertainties of the order of a few in the determination of the mass of the observed galaxies, due to uncertainties in their distance, the bolometric luminosity and the light-to-mass ratio. The majority of the models presented here produce the observed number of LMXBs to within a factor of 3, consistent with the galaxy mass uncertainties. Exceptions are models with transient treatment E or F (see Table \\ref{models_tran}) and models with high CE efficiencies (models 21-28, see \\ref{models}) which greatly overproduce hight luminosity LMXBs. Furthermore small changes in the CE efficiency can change the total number of sources produced by a model by a factor of two, without changing significantly the shape of the XLF (compare for example models $\\rm 10A^{IT}$, $\\rm 14A^{IT}$, $\\rm 18A^{IT}$ in the online supplemental material). In view of these uncertainties and our limited parameter space exploration for the models, we consider this normalization agreement satisfactory, but do not use it as an actual constraint on the models. We do not claim that the work presented in this Paper is a complete PS study of field LMXBs in elliptical galaxies. It is meant to be a first effort in interpreting the recent deep \\emph{Chandra} observations of the two elliptical galaxies NGC3379 and NGC4278. Throughout the Paper we identify the caveats of our analysis, and we intend to address them in our future work. It turns out that the realistic treatment of the outburst properties of transient LMXBs is crucial for the modeling of XLFs of extragalactic populations. The derivation of an empirical correlation between the outburst luminosity and the period of Galactic NS-LMXBs, similar to the one derived by \\citet{PDM2005} for Galactic BH-LMXBs, will provide a better understanding of the differences in the transient behavior of these two classes of X-ray sources. Another simplifying assumption we made, is the constant outburst luminosity of transient LMXBs. The use of model lightcurves for the outburst phases of transient XLFs will possibly affect the shape of the modeled XLFs. We note that most of our models produce many systems with X-ray luminosity below the observational limit and down to $10^{35}\\rm \\, erg \\, s^{-1}$. The integrated diffuse X-ray emission from these galaxies can put a strong constraint on our models. The total luminosity of the observed diffuse emission will also include emission from gas and stellar coronae \\citep{PF1994, Revnivtsev2007} and thus must be higher than the integrated luminosity of all the LMXBs in our models with luminosities below the observational detection limit. \\clearpage \\newpage" }, "0801/0801.1587_arXiv.txt": { "abstract": "We present a comparison between the 2001 XMM-Newton and 2005 Suzaku observations of the quasar, PG\\,1211+143 at $z=0.0809$. Variability is observed in the 7 keV iron K-shell absorption line (at 7.6 keV in the quasar frame), which is significantly weaker in 2005 than during the 2001 XMM-Newton observation. From a recombination timescale of $<4$ years, this implies an absorber density $n>4\\times10^{3}$\\,cm$^{-3}$, while the absorber column is $5\\times10^{22} 10^{16}$~cm$^{-2}$, equivalent to Hydrogen columns of $N_H\\sim 10^{19}$~cm$^{-2}$. Much more controversial was the potential association of substantially higher columns of gas to such local material. McKernan et al (2004; 2005) noticed that several AGN with strong He-- or H--like Fe absorption features had these lines at energies which were approximately consistent with the local standard of rest. For any single object the approximate match between the putative blueshifted outflow velocity and the galaxy redshift could be coincidental, but McKernan et al (2004; 2005) pointed out three AGN showing this trend, with redshifts spanning $\\sim 0.008-0.15$ (MCG--6--30--15, PG1211+143 and PDS~456). However, the derived columns of $N_H\\sim 10^{23}$~cm$^{-2}$ are high even for an intrinsic AGN outflow (Pounds et al 2003, hereafter P03; Reeves et al 2003). If these are instead of local origin then it requires a tremendously significant change to our understanding of the Galactic halo environment (McKernan et al 2004; 2005). Here we use new Suzaku data to show that the highly ionised Fe absorption in PG~1211+143 is variable on a timescale of years. This conclusively demonstrates that it is intrinsic to the AGN, and not associated with our Galaxy or Local Group. We collate recent results to show that this is also the case for most other AGN with strong Fe K absorption lines, showing that powerful outflows are associated with luminous accretion flows. ", "conclusions": "We show that the observed variability in the absorption line in PG~1211+143 between the XMM-Newton and Suzaku data taken 4 years apart conclusively rules out a diffuse gas origin such as the local Galactic halo or WHIM. The gas must be associated with the AGN, as is further evidenced by its large column density in the XMM-Newton 2001 data which is far too high for local or intergalactic gas. The conclusive identification with the AGN, and the implausibility of any alternative line transition other than iron K$\\alpha$ means that the large outflow velocity of $\\sim 0.1$c is inescapable. Such material is predicted from the winds which are produced from luminous accretion discs in AGN (Proga \\& Kallman 2004). The coincidence of the outflow velocity with the source redshift in these data is not significant. Other powerful AGN which show iron K absorption features clearly show a range of (intrinsic and observed) velocities, removing the apparent trend for the line energy to appear at the rest energy for these transitions. Thus it is clear that this material is a mildly relativistic outflow from the AGN. Its high velocity means that its kinetic energy can be comparable to the bolometric radiated luminosity of the AGN (King \\& Pounds 2003; Pounds et al 2003; Pounds \\& Reeves 2007), yet its high ionisation means that it is effectively invisible in all other wavebands. While such a major component of AGN energetics could go unnoticed in individual objects, there is increasing evidence that strong AGN feedback controls galaxy formation and evolution. Mildly relativistic winds provide more efficient heating than the jet due to their impact on a larger area, so these highly ionised disc winds may be the key to understanding the growth of structure in the Universe." }, "0801/0801.1064_arXiv.txt": { "abstract": "We briefly review capabilities and requirements for future instrumentation in UV- and X-ray astronomy that can contribute to advancing our understanding of the diffuse, highly ionised intergalactic medium. ", "introduction": "\\label{Introduction} A convergence of recent theoretical and observational work has generated rapidly growing interest in the physics of a highly ionised intergalactic medium (IGM). At low redshifts, this phase of the IGM likely holds the balance of the baryon mass density not accounted for by the mass densities in stars, diffuse gas in galaxies, the local Ly$\\alpha$ forest, the Intracluster Medium (ICM) in galaxy clusters and groups, etc. \\citep{FP04}. Model calculations suggest that in fact the baryon density in the IGM could indeed be comparable to the summed baryon densities of the known mass components, as would be required to reach the baryon density determined using independent arguments (big bang nucleosynthesis and the measured abundances of the light elements, and the measured fluctuation spectrum of the cosmic microwave background). It is clear that the study of non-equilibrium phenomena in the low-redshift diffuse intergalactic medium, both within and outside bound structures, can produce a wealth of information on a large range of important astrophysical problems, ranging from the formation of galaxies to the generation of the first dynamically important magnetic fields. In this paper, we will attempt to summarise the requirements and prospects for significant progress on the observational study of these phenomena, considered in the light of planned or proposed instrumentation. We begin with a slightly abstract discussion of required and desired instrumental capabilities (spectral resolution, sensitivity, etc.), before we consider specific future or proposed instruments. ", "conclusions": "" }, "0801/0801.1778_arXiv.txt": { "abstract": "{} {Investigation of the dense gas, the outflows and the continuum emission from the massive twin cores NGC6334I and I(N) at high spatial resolution.} {We imaged the region with the Australia Telescope Compact Array (ATCA) at 3.4\\,mm wavelength in continuum as well as CH$_3$CN$(5_K-4_K)$ and HCN(1--0) spectral line emission.} {While the continuum emission in NGC6334I mainly traces the UCH{\\sc ii} region, toward NGC6334I(N) we detect line emission from four of the previously identified dust continuum condensations that are of protostellar or pre-stellar nature. The CH$_3$CN$(5_K-4_K)$ lines are detected in all $K$-components up to energies of 128\\,K above ground toward two protostellar condensations in both regions. We find line-width increasing with increasing $K$ for all sources, which indicates a higher degree of internal motions of the hotter gas probed by these high K-transitions. Toward the main mm and CH$_3$CN source in NGC6334I we identify a velocity gradient approximately perpendicular to the large-scale molecular outflow. This may be interpreted as a signature of an accretion disk, although other scenarios, e.g., an unresolved double source, could produce a similar signature as well. No comparable signature is found toward any of the other sources. HCN does not trace the dense gas well in this region but it is dominated by the molecular outflows. While the outflow in NGC6334I exhibits a normal Hubble-law like velocity structure, the data are consistent with a precessing outflow close to the plane of the sky for NGC6334I(N). Furthermore, we observe a wide ($\\sim$15.4\\,km\\,s$^{-1}$) HCN absorption line, much broader than the previously observed CH$_3$OH and NH$_3$ absorption lines. Several explanations for the difference are discussed.} {} ", "introduction": "The massive twin cores NGC6334I and I(N) at a distance of 1.7\\,kpc in the southern hemisphere \\citep{neckel1978,straw1989} have been subjected to investigations for more than two decades. The two regions are located at the north-eastern end of the much larger molecular cloud/H{\\sc ii} region complex NGC6334 (e.g., \\citealt{rodriguez1982,gezari1982,depree1995,kraemer1999b,sandell2000,carral2002}). The reason why NGC6334I (synonymous with NGC6334F) and NGC6334I(N) are so interesting from a comparison point of view is that they are only separated by approximately 1\\,parsec, hence they share a similar large-scale molecular environment, but they exhibit extremely different characteristics likely because they are at different evolutionary stages. Both regions have been studied in much detail over the last decades; recent summaries of the past observations can be found, e.g., in \\citet{hunter2006}, \\citet{beuther2007b} or \\citet{rodriguez2007}. Here we just outline their main characteristics. NGC6334I is a prototypical hot molecular core right at the head of a cometary ultracompact H{\\sc ii} (UCH{\\sc ii}) region \\citep{depree1995,kraemer1995}. It exhibits rich spectral line emission \\citep{mccutcheon2000,thorwirth2003,schilke2006}, a bipolar outflow \\citep{bachiller1990,leurini2006} and H$_2$O, OH, CH$_3$OH class {\\sc ii} and NH$_3$(3,3)/(6,6)/(8,6)/(11,9) maser emission \\citep{moran1980,forster1989,gaume1987,brooks2001,norris1993,caswell1997,walsh1998,beuther2007b,walsh2007}. In contrast to that, up to very recently NGC6334I(N) was considered a typical cold core since no mid-infrared and only faint near-infrared emission was detected \\citep{gezari1982,tapia1996,persi2005}. Furthermore, weak cm continuum and class {\\sc i} and {\\sc ii} CH$_3$OH maser emission was reported \\citep{carral2002,kogan1998,caswell1997,walsh1998}. The spectral line forest is considerably less dense compared to NGC6334I \\citep{thorwirth2003}, however, a few species are stronger toward NGC6334I(N) (\\citealt{sollins2004c}, Walsh et al.~in prep.). In addition, \\citet{megeath1999} report the detection of a molecular outflow in this region as well. In summary, both regions show signs of active star formation, however, the southern region NGC6334I appears to be in a more advanced evolutionary stage than the northern region NGC6334I(N). To better characterize this intriguing pair of massive star-forming regions, we started a concerted campaign from cm to mm wavelengths with the Australia Telescope Compact Array (ATCA), the Submillimeter Array (SMA) and the Mopra single-dish telescope. The previous ATCA NH$_3$(1,1) to (6,6) line observations revealed compact warm gas emission from both regions \\citep{beuther2005e,beuther2007b}, and temperatures estimated to exceed 100\\,K. While toward NGC6334I(N) the low energy NH$_3$ lines showed only extended emission, the high energy lines finally revealed compact gas components. The NH$_3$(6,6) line profile from NGC6334I(N) allowed speculation about a potential accretion disk. CH$_3$OH was strong in absorption toward the southern UCH{\\sc ii} region in NGC6334I, indicative of expanding gas. In the mm continuum emission, \\citet{hunter2006} used the SMA to resolve several mm continuum sources toward both regions (4 in NGC6334I and 7 in NGC6334I(N)). Furthermore, Hunter et al.~(in prep.) identified an additional SiO outflow in NGC6334I(N) that has its orientation in north-east south-west direction, approximately perpendicular to the one previously reported by \\citet{megeath1999}. Here we present 3.4\\,mm continuum and HCN/CH$_3$CN spectral line observations obtained with the new 3\\,mm facility at the ATCA. These observations shed light on the outflow and dense gas properties of both regions as well as on the continuum emission from the embedded protostellar objects and the UCH{\\sc ii} region. ", "conclusions": "Our 3.4\\,mm continuum and CH$_3$CN/HCN spectral line study of the massive twin cores NGC6334I and I(N) reveals many new insights into that intriguing pair of massive star-forming regions. Both sets of spectral lines as well as the continuum emission are clearly detected toward both targets. While the continuum emission in NGC6334I mainly follows the UCH{\\sc ii} regions and the strongest protostellar 1.4\\,mm peak is detected only at a $\\sim 3\\sigma$ level, in NGC6334I(N), the 3.4\\,mm continuum emission traces four of the previously identified protostellar or pre-stellar condensations. In both regions, the whole CH$_3$CN$(5_K-4_K)$ $K$-ladder from $K=0$ to 4 is detected toward the strongest protostellar condensations. While the emission is in most cases so optically thick that temperature estimates are prohibited, toward the secondary mm peak in NGC6334I(N) we can estimate a temperature of $170\\pm50$\\,K. Toward all four detected CH$_3$CN emission sources, we find a correlation between increasing line-width and increasing excitation temperature of the $K$ components. Since increasing excitation temperatures are expected closer to the protostars, this implies more internal motions, e.g., outflow, infall or rotation, the closer one gets to the central protostar. Similar signatures are observed in the CH$_3$CN 2nd moment maps. To investigate potential rotation, we produced 1st moment maps toward all CH$_3$CN peaks, and fitted the channel peak positions toward mm1 in NGC6334I to effectively increase the spatial resolution. We identify a velocity gradient toward mm1 in NGC6334I that is oriented approximately perpendicular to the known large-scale outflow. This may be interpreted as a signature of a rotating structure, maybe associated with a massive accretion disk. While early rotation-disk claims for that region were on scales of the molecular outflow \\citep{jackson1988}, we are now reaching spatial scales of the order a few hundred AU, much more reasonable for accretion disks (e.g., \\citealt{yorke2002,krumholz2006b}). However, we have to stress that the rotation signatures are not conclusive yet because, e.g., an unresolved double-source could produce similar signatures. Further investigation at higher angular resolution are required to resolve this issue. While higher angular resolution images with the ATCA at 3\\,mm wavelengths are difficult because of decreasing phase stability with increasing baseline length, one may tackle that problem in some highly excited NH$_3$ lines in the 12\\,mm band. Furthermore, ALMA will allow the investigation of this source in much greater depth. Toward the previously suggested disk candidate in NGC6334I(N) we cannot identify a similar velocity gradient. In contrast to conventional wisdom that HCN traces the dense gas cores, we find it most prominently in the molecular outflows of both massive star-forming regions. The velocity structure of the outflow in NGC6334I is relatively normal and follows the well-known Hubble-law for molecular outflows. In addition to that, we find the broadest HCN line-width toward the main mm continuum peak mm1. In contrast to the previously found elongation of NH$_3$(6,6) maser emission indicating that mm2 could drive the molecular outflow, these data suggest that mm1 harbors the outflow driving source. However, it is also possible that there are two molecular outflows with different properties that are preferentially detected in different tracers (HCN in this work and CO for the previous larger-scale outflow detection at a slightly different position-angle). The velocity structure of the NGC6334I(N) outflow is more peculiar. There we find a change between blue- and red-shifted outflow emission on one side of the outflow. Taking into account the additionally bended morphology of that outflow, a possible explanation for this velocity structure is a precessing outflow close to the plane of the sky. Furthermore, HCN exhibits a broad absorption feature with a line-width of $\\sim 15.4$\\,km\\,s$^{-1}$ toward the UCH{\\sc ii} region in NGC6334I. Comparing the line-width of the previously observed absorption features of NH$_3$ and CH$_3$OH, which are of the order 2\\,km\\,s$^{-1}$, with the HCN line-width as well as the line-width observed in the ionized gas of $\\sim$32\\,km\\,s$^{-1}$, two explanations are possible to explain the different line-widths. The velocity gradient identified in the ionized gas indicates that it may be influenced by the molecular outflow close by. If the opacity of HCN were very large it traced only the outer gas layers around the UCH{\\sc ii} region and could also be influenced by the outflow. On the other hand, NH$_3$ is optically thin based on the absent absorption in the hyperfine satellite lines. If HCN were optically thin as well, another possible explanation would be based on the different critical densities of the various molecules: In the picture of an expanding UCH{\\sc ii} region, the densities close to the UCH{\\sc ii} region surface should be higher than those further outside. Hence HCN may trace the gas closer to the expanding UCH{\\sc ii} region and is then much stronger affected by the expansion process that the lower-density regions further out that are traced by NH$_3$ and CH$_3$OH. To differentiate between both models, observations of a rarer HCN isotopologues are required to derive its optical depth." }, "0801/0801.3243_arXiv.txt": { "abstract": "The Alpha Magnetic Spectrometer (AMS), whose final version AMS-02 is to be installed on the International Space Station (ISS) for at least 3 years, is a detector designed to measure charged cosmic ray spectra with energies up to the TeV region and with high energy photon detection capability up to a few hundred GeV, using state-of-the art particle identification techniques. Among several detector subsystems, AMS includes a proximity focusing RICH enabling precise measurements of particle electric charge and velocity. The combination of both these measurements together with the particle rigidity measured on the silicon tracker endows a reliable measurement of the particle mass. The main topics of the AMS-02 physics program include detailed measurements of the nuclear component of the cosmic-ray spectrum and the search for indirect signatures of dark matter. Mass separation of singly charged particles, and in particular the separation of deuterons and antideuterons from massive backgrounds of protons and antiprotons respectively, is essential in this context. Detailed Monte Carlo simulations of AMS-02 have been used to evaluate the detector's performance for mass separation at different energies. The obtained results and physics prospects are presented. ", "introduction": "The Alpha Magnetic Spectrometer (AMS)\\cite{bib:ams}, whose final version AMS-02 is to be installed on the International Space Station (ISS) for at least 3 years, is a detector designed to study the cosmic ray flux by direct detection of particles above the Earth's atmosphere using state-of-the-art particle identification techniques. AMS-02 is equipped with a superconducting magnet cooled by superfluid helium. The spectrometer is composed of several subdetectors: a Transition Radiation Detector (TRD), a Time-of-Flight (TOF) detector, a Silicon Tracker, Anticoincidence Counters (ACC), a Ring Imaging \\CK\\ (RICH) detector and an Electromagnetic Calorimeter (ECAL). Fig.~\\ref{amsdet} shows a schematic view of the full AMS-02 detector. A preliminary version of the detector, AMS-01, was successfully flown aboard the US space shuttle Discovery in June 1998\\cite{bib:ams01res}. \\begin{figure}[htb] \\center \\vspace{0.2cm} \\mbox{\\epsfig{file=ams2.eps,width=0.48\\textwidth,clip=}} \\caption{Exploded view of the AMS-02 detector.\\label{amsdet}} \\vspace{-0.5cm} \\end{figure} The main goals of the AMS-02 experiment are: \\begin{itemize} \\item A precise measurement of charged cosmic ray spectra in the rigidity region between \\mbox{$\\sim$ 0.5 GV} and \\mbox{$\\sim$ 2 TV}, and the detection of photons with energies up to a few hundred GeV; \\item A search for heavy antinuclei ($Z \\ge$ 2), which if discovered would signal the existence of cosmological antimatter; \\item A search for dark matter constituents by examining possible signatures of their presence in the cosmic ray spectrum. \\end{itemize} The long exposure time and large acceptance (0.5 m${}^2\\cdot$sr) of AMS-02 will enable it to collect an unprecedented statistics of more than $10^{10}$ nuclei. ", "conclusions": "AMS-02 will provide a major improvement on the current knowledge of cosmic rays. A total statistics of more than 10${}^{10}$ events will be collected during its operation. Detailed simulations have been performed to evaluate the detector's particle identification capabilities, in particular those of the RICH. Simulation results show that the separation of light isotopes is feasible. Using a set of simple cuts based on event data, relative mass resolutions of $\\sim$~2 \\% and rejection factors higher than 10${}^4$ have been attained in D/p separation at energies of a few GeV/nucleon. The separation procedure presented here might be crucial for the identification of an antideuteron flux resulting from neutralino annihilation." }, "0801/0801.1070_arXiv.txt": { "abstract": "We use an independent new sample of cataclysmic variables (CVs), constructed by selecting objects for H$\\alpha$ emission, to constrain the properties of the intrinsic CV population. This sample is restricted to systems that are likely to be non-magnetic and unevolved; it consists of 17 CVs, of which at least 10 have orbital periods above 3~h. We find that even very generous allowance for selection effects is not sufficient to reconcile the large ratio of short- to long-period CVs predicted by standard CV evolution theory with the observed sample, possibly implying that short-period systems evolve faster than predicted by the disrupted magnetic braking model. This would require that an angular momentum loss mechanism, besides gravitational radiation, acts on CVs with orbital periods below the period gap. To bring the model into agreement with observations, the rate of angular momentum loss below the period gap must be increased by a factor of at least 3, unless the model also over estimates the angular momentum loss rate of long-period CVs. ", "introduction": "Despite its wide relevance, the evolution of cataclysmic variables (CVs) is still not very well understood. Much of the problem stems from the fact that strong selection effects act on observed samples of CVs, making it difficult to constrain theory observationally. The angular momentum loss rate ($-\\dot{J}$) is the crucial ingredient of CV evolution theory. Angular momentum loss leads to mass transfer from the secondary to the white dwarf. The relation between the mass transfer rate ($\\dot{M}$) and $\\dot{J}$ depends on the structure of the secondary star, and the reaction of the secondary to mass loss determines the orbital period ($P_{orb}$) evolution of the system. Population synthesis methods combine the changing $\\dot{M}$ and $P_{orb}$ with a model of the birthrate of CVs to predict the present-day distribution of CVs as a function of $M_1$, $M_2$ (the mass of the white dwarf and secondary, respectively), $\\dot{M}$, and $P_{orb}$. Only one of these parameters, $P_{orb}$, can practically be measured for a large number of CVs (but see \\citealt{Patterson98} for an indirect method to measure mass ratios). The orbital period distribution of known CVs is therefore one of the few observational properties of the CV population that can be used to constrain theory. The other obvious constraint is the observed CV space density. In comparing both the size and period distribution of the known CV population to theoretical predictions, observational bias must be accounted for. This requires that observed CV samples be restricted to objects selected in well-defined (and preferably homogeneous) ways. A long-standing problem is that theory predicts a large population of short-period CVs, consisting mostly of period bouncers (systems that have evolved beyond the period minimum at about 76~min), which is not observed. The models of \\cite{Kolb93} and \\cite{HowellRappaportPolitano97} both predict that only $\\simeq 1$\\% of CVs are long-period systems, while roughly 70\\% are period bouncers. Existing observations have already been used to argue that the short-period CV population cannot be as large as predicted (e.g. \\citealt{Patterson98}; \\citealt{PretoriusKniggeKolb07}). We will do the same here, using a new and independent CV sample. The CV samples that have been available to date are heavily biased against the intrinsically faint, short-period CVs, mainly because most have bright flux limits. The new sample considered here is also limited to apparently bright systems. However, almost all surveys incorporate a second selection cut (most commonly a blue cut) that also discriminates against the discovery of short-period CVs. The CV sample that we have constructed differs from most existing samples in that the only selection criterion (other than a flux limit) is based on line emission. The spectra of the majority of CVs show Balmer emission lines, originating mainly in the accretion flow. The luminosity of CVs is anti-correlated to the equivalent widths (EWs) of their emission lines (\\citealt{Patterson84}; \\citealt{WithamKniggeGansicke06}; we will take EWs of emission lines as positive throughout). The explanation for this anti-correlation is that intrinsically faint CVs are low-$\\dot{M}$ systems with low density discs, in which recombination line cooling is very efficient. Therefore, in contrast to other selection criteria (such as blue optical colours and variability), an emission line EW-based selection cut should favour the discovery of intrinsically faint, short-period systems. We have used the AAO/UKST SuperCOSMOS H$\\alpha$ Survey (SHS) to define a homogeneous sample of CVs, selected on the basis of H$\\alpha$ emission. Observations of the new CVs were presented in an earlier paper (\\citealt{HalphaI}; hereafter Paper I). Here we describe the construction and completeness of the sample, examine the observational biases affecting it, and compare it to the predictions of theory. ", "conclusions": "We have compared a homogeneous CV sample, selected for H$\\alpha$ emission, to a model CV population based on standard CV evolution theory. The magnitude limit and Galactic latitude range of the observed sample was modelled in some detail, while conservative assumptions were made to account for the effects of variability and the $\\mathrm{EW}(\\mathrm{H}\\alpha)$-based selection cut. The model population is inconsistent with the observed sample. Specifically, the model predicts relatively too many short-period CVs. This confirms earlier results, based on independent observations. The reason for the mismatch between the predicted and observed ratio of short- to long-period CVs may be that the theoretical evolutionary time-scale for CVs below the period gap is too long. A (very simplistic) consideration of the relative numbers of long- and short-period CVs included in the sample indicates that the disrupted magnetic braking model underestimates $-\\dot{J}$ of short-period CVs by a factor of at least 3, assuming that the model is correct for long-period CVs. Although surveys now in progress will in future provide much better observational constraints on CV evolution theory than can be derived at the moment, it is already clear that the standard magnetic braking model is in need of revision. Furthermore, it seems that the correct approach to take is to investigate angular momentum loss rates in excess of the gravitational radiation rate in CVs below the period gap." }, "0801/0801.1841_arXiv.txt": { "abstract": "We present refined values for the physical parameters of transiting exoplanets, based on a self-consistent and uniform analysis of transit light curves and the observable properties of the host stars. Previously it has been difficult to interpret the ensemble properties of transiting exoplanets, because of the widely different methodologies that have been applied in individual cases. Furthermore, previous studies often ignored an important constraint on the mean stellar density that can be derived directly from the light curve. The main contributions of this work are \\emph{i}) a critical compilation and error assessment of all reported values for the effective temperature and metallicity of the host stars; \\emph{ii}) the application of a consistent methodology and treatment of errors in modeling the transit light curves; and \\emph{iii}) more accurate estimates of the stellar mass and radius based on stellar evolution models, incorporating the photometric constraint on the stellar density. We use our results to revisit some previously proposed patterns and correlations within the ensemble. We confirm the mass-period correlation, and we find evidence for a new pattern within the scatter about this correlation: planets around metal-poor stars are more massive than those around metal-rich stars at a given orbital period. Likewise, we confirm the proposed dichotomy of planets according to their Safronov number, and we find evidence that the systems with small Safronov numbers are more metal-rich on average. Finally, we confirm the trend that led to the suggestion that higher-metallicity stars harbor planets with a greater heavy-element content. ", "introduction": "\\label{sec:introduction} The transiting exoplanets are only a small subset of all the known planets orbiting other stars, but they hold tremendous promise for deepening our understanding of planetary formation, structure, and evolution. Observations of transits and occultations\\footnote{The word \\emph{transit} is sometimes assumed in the field of exoplanet research to be synonymous with \\emph{eclipse}. In reality, it has a more restricted meaning and has long been used in the eclipsing binary field to describe an eclipse of the larger object by the smaller one. The term \\emph{occultation} is used to refer to the passage of the smaller object (in this case, the planet) behind the larger one (the star) \\citep[see, e.g.,][]{Popper:76}. To avoid confusion, we advocate that \\emph{occultation} or \\emph{secondary eclipse} are preferable to neologisms such as ``secondary transit'' or ``anti-transit''.} (along with the spectroscopic orbit of the host star) not only allow one to measure the mass and radius of the planet, but also provide opportunities to measure the stellar spin-orbit alignment \\citep{Queloz:00, Winn:07a}, the planetary brightness temperature \\citep{Charbonneau:05, Deming:05}, the planetary day-night temperature difference \\citep{Knutson:07}, and even absorption lines of planetary atmospheric constituents \\citep{Charbonneau:02, Vidal-Madjar:04, Tinetti:07}. These and other observations have been accompanied by theoretical progress in modeling the physical processes in the planetary interiors and atmospheres, as well as the planets' interactions with their parent stars. This rapid progress has been stimulated in no small measure by the remarkable diversity of planet characteristics that has been found among the members of the transiting ensemble. The accuracy and precision with which the properties of the planet can be derived from transit data depend strongly on whatever measurements and assumptions are made regarding the host star. For example, for a given value of the transit depth, the inferred planetary radius scales in proportion to the assumed stellar radius; and for a given spectroscopic orbit of the host star, the inferred planetary mass scales as the two-thirds power of the stellar mass. In the literature on transiting planets, a wide variety of methods have been used to estimate the radius and mass of the parent star, ranging from simply looking them up in a table of average stellar properties as a function of spectral type, all the way to fitting detailed stellar evolutionary models constrained by the luminosity, effective temperature, and other observations that may be available for the star. As a result, the ensemble of planet properties at our disposal is inhomogeneous, and in many cases the uncertainty of those determinations is dominated by systematic errors in the stellar parameters that are treated differently by different investigators. This situation is unfortunate because it hinders our ability to gauge the reliability of any patterns that are discerned among the ensemble properties of transiting exoplanets. With 23 systems that have been reported in the literature, this subfield should be poised for the transition from a handful of results to a large and diverse enough sample for meaningful general conclusions to be drawn, but the heterogeneity of reported results is clearly an obstacle. It may be surprising that we are limited in many cases by our knowledge of the properties of the parent stars; one would think that stellar physics is a ``solved problem'' in comparison to exoplanetary physics. However it must be remembered that the host stars are usually isolated (and therefore no dynamical mass measurement is possible), and that many of the host stars are distant enough that they do not even have measured parallaxes. Of the 23 cases in the literature, five have \\hip\\ parallaxes \\citep{Perryman:97}. For those few it is fairly straightforward to estimate the stellar properties, but for the other systems, more indirect methods have been used. These indirect methods often rely on the value of the stellar surface gravity that is derived by measuring the depths and shapes of gravity-sensitive absorption lines in the stellar spectrum. This is a notoriously difficult measurement and the result is often strongly correlated with other parameters that affect the spectrum. Recently, however, \\cite{Sozzetti:07}, Holman et al.~(2007), and others demonstrated that it is possible to do better by using the information about the mean stellar density that is encoded in the transit light curve. This information was typically overlooked prior to these studies. The study presented in this paper was motivated by the desire for a homogeneous analysis, and by the desire to take advantage of the photometric estimates of the stellar mean density. We have revisited the determination of the stellar parameters for all of the transiting planets that have been reported in the literature. We have taken the opportunity to merge all existing measurements of the atmospheric parameters (mainly the effective temperature and metallicity) with the goal of presenting the best possible values. We have chosen a uniform method for analyzing photometric data and have re-analyzed existing light curves where necessary to provide homogeneity. Our hope was that by applying these procedures across the board, we and other investigators could view the ensemble properties with greater clarity and uncover any interesting clues the transiting planets might provide us about the origin, structure, and evolution of exoplanets. This paper is organized as follows. \\S\\,\\ref{sec:stellar} describes the procedures by which we estimated the stellar properties, using the available spectroscopic and photometric datasets, and a particular set of theoretical stellar evolution models. As a check on the models, \\S\\,\\ref{sec:modchecks} compares the results of a subset of our calculations with those derived from a different set of evolutionary models. \\S\\,\\ref{sec:obschecks} investigates alternate ways of estimating the stellar properties. \\S\\,\\ref{sec:gj436} deals with GJ~436, which needs special treatment because the host star has such a lower mass than the other host stars. \\S\\,\\ref{sec:results} presents the final results for the planetary parameters. \\S\\,\\ref{sec:discussion} uses the new results to check on some of the previously proposed correlations among the properties of the transiting ensemble, and \\S\\,\\ref{sec:remarks} provides final remarks. The Appendix lists the data sets and other issues that are particular to each system. ", "conclusions": "\\label{sec:discussion} Many investigators have sought and claimed possible correlations between various stellar and planetary parameters of transiting systems. In principle such correlations could lead to important insights into the formation, structure, and evolution of exoplanets. A primary motivation for presenting a more complete, accurate, and homogeneous set of these parameters in this work was to facilitate such studies. The relatively large array of properties now available offers the opportunity to find new correlations, or to revisit old ones incorporating additional variables. While it is beyond the scope of the present work to investigate all possible correlations with statistical rigor, in this section we check on three of the most intriguing and potentially important relations that have been proposed. \\subsection{Planetary mass versus orbital period} \\label{sec:massperiod} \\cite{Mazeh:05} were the first to point out the apparent correlation between $M_p$ and $P$ for transiting planets \\citep[see also][]{Gaudi:05}. The original suggestion was based on only 6 systems, but additional discoveries have generally supported the trend of decreasing mass with longer periods, although the scatter has also become larger. This is shown in Figure~\\ref{fig:mp}a. HAT-P-2b would be an extreme outlier in this plot; we have excluded it because it is so much more massive than the other planets and may belong to a different category of planet (see \\S\\,\\ref{sec:safronov}). Similarly, we have excluded GJ~436b because it is so much less massive than the others, and may be a rocky or rock-ice planet rather than a gas giant. A simple linear fit is shown for reference (dashed line). \\begin{figure} % \\vskip 0.0in \\epsscale{1.3} {\\hskip -0.2in\\plotone{f9xv.ps}} \\vskip 0.0in \\figcaption[]{(a) $M_p$ as a function of period for all transiting planets except HAT-P-2b ($M_p = 8.7$~M$_{\\rm Jup}$, $P = 5.63$~d) and GJ~436b ($M_p = 0.073$~M$_{\\rm Jup}$, $P = 2.64$~d); see text. The dashed line is a linear fit. (b) $O\\!-\\!C$ residuals $\\Delta M_p$ from the linear fit in the top panel shown as a function of the metallicity of the host star. The dashed line is a linear fit. (c) Same as (a), with the dependence on [Fe/H] removed based on the fit in (b). The fitted line has the expression $M_p = (+1.70 \\pm 0.13) - (0.281 \\pm 0.044) \\times P$. \\label{fig:mp}} \\end{figure} We also investigated the scatter in this relation, seeking any ``third variable'' that might correlate with the residuals. We seem to have found such a third variable: the metallicity of the host star. Figure~\\ref{fig:mp}b displays the $O\\!-\\!C$ residuals from the top panel as a function of [Fe/H], indicating a rather clear correlation (dashed line): $\\Delta M_p = (+0.152 \\pm 0.050) - (1.17 \\pm 0.23) \\times {\\rm [Fe/H]}$. After removal of this trend, the relation between $M_p$ and period becomes tighter (Figure~\\ref{fig:mp}c). The scatter in the mass-period relation is reduced from 0.26~M$_{\\rm Jup}$ in the top panel to 0.17~M$_{\\rm Jup}$ in the bottom panel. It seems unlikely that this is a statistical fluke. However, the scatter is still larger than the formal observational uncertainties, suggesting that these three variables are not completely determinative. What might be the implications of this metallicity dependence? It has been proposed that the mass-period relation is related to the process by which close-in exoplanets migrated inward from their formation sites, or more specifically, to the mechanism that halts migration at orbital periods of a few days. The trend of larger masses at shorter orbital periods could suggest that the halting mechanism depends on mass, and larger planets are able to migrate further in. The dependence on metallicity may then be interpreted to indicate that planets in metal-poor systems need to be more massive in order to migrate inward to the same orbital period as more metal-rich planets. In this context, the trend in Figure~\\ref{fig:mp}b could be interpreted as evidence that the efficiency of the migration (or halting) mechanism is affected to some degree by the chemical composition. A dependence of migration on metallicity is in fact predicted by some theories, and could arise, as pointed out by \\cite{Sozzetti:06}, either from slower migration rates in metal-poor protoplanetary disks \\citep{Livio:03, Boss:05} or through longer timescales for giant planet formation around metal-poor stars, which would effectively reduce the efficiency of migration before the disk dissipates \\citep{Ida:04, Alibert:05}. However, the above processes are more aimed at addressing the apparent lack of short-period planets among very metal-poor stars claimed by some authors \\citep[see][]{Sozzetti:06}, whereas among the transiting planets it is not the lack of more metal-poor examples we are concerned with, but rather their different properties (such as mass) compared to metal-rich planets \\emph{at the same orbital period}. To give a quantitative example, we find that for a period of 2.5 days (near the average for known transiting systems), the mass of a planet with [Fe/H] $= -0.2$ is $\\sim$40\\% larger than one with average metallicity ([Fe/H] = +0.13, excluding HAT-P-2b and GJ~436b), while the mass of a planet with [Fe/H] $= +0.4$ is about 30\\% smaller. This range of metallicities, from [Fe/H] $= -0.2$ to +0.4 (covering a factor of 4 in metal enhancement), is approximately the full range observed. We note that the strength of the metallicity effect is modest rather than overwhelming, and this too is in agreement with theoretical expectations \\citep[e.g.,][]{Livio:03}. The massive planet HAT-P-2b obviously does not conform to this trend of $M_p$ versus period, which may indicate some fundamental difference either in its formation or migration. An alternative interpretation, also proposed by \\cite{Mazeh:05}, is that the $M_p$ versus $P$ relation is more a reflection of survival requirements in close proximity to the star, due to thermal evaporation from the extreme UV flux. Close-in planets must be more massive to avoid ablation to the point of undetectability. The role of metallicity in this case would be through the difference in the internal structure \\citep{Santos:06}. Metal-rich planets have been suggested to be more likely to develop rocky cores \\citep[][see also \\S\\,\\ref{sec:cores}]{Pollack:96, Guillot:06, Burrows:07}. If the presence of such a core somehow slows down or prevents complete evaporation, as has been proposed \\citep{Baraffe:04, Lecavelier:04}, survival at a given period would then have a dependence on metallicity. \\begin{figure} \\vskip -0.35in \\epsscale{1.3} {\\hskip -0.2in\\plotone{f10.eps}} \\vskip -0.3in \\figcaption[]{Surface gravity versus orbital period for all transiting planets except HAT-P-2, which is off the scale ($g_p \\sim 234$ m~s$^{-2}$). A linear fit is shown.\\label{fig:loggperiod}} \\end{figure} A diagram related to the one considered above is that of planetary surface gravity versus orbital period, first presented by \\cite{Southworth:07}. Because $g_p$ does not require knowledge of the mass or radius of the star, it is in a sense a cleaner quantity that should be free from systematic errors in $M_{\\star}$ and $R_{\\star}$ and is nearly independent of stellar evolution models. This not true of other bulk properties such as the mean planet density. An updated version of the $g_p$ versus $P$ relation is shown in Figure~\\ref{fig:loggperiod}, along with a linear fit. We examined the residuals from this linear fit for a possible correlation with stellar metallicity, as we did earlier for the case of the $M_p$ versus $P$ diagram, but we found none. Given that $g_p \\propto M_p/R_p^2$, we note that the apparent influence of metallicity on the planetary radii must be contributing significantly to the scatter in the $g_p$ versus period relation. We discuss this further in \\S\\,\\ref{sec:cores}. \\subsection{Safronov number versus equilibrium temperature} \\label{sec:safronov} Recently, \\cite{Hansen:07} proposed a distinction between two classes of hot Jupiters, based on a consideration of the Safronov number and the zero-albedo equilibrium temperature. The Safronov number is a measure of the ability of a planet to gravitationally scatter other bodies \\citep{Safronov:72}, and is defined as $\\Theta = \\frac{1}{2}(V_{\\rm esc}/V_{\\rm orb})^2 = (a/ R_p)(M_p/ M_{\\star})$, the ratio between the escape velocity and the orbital velocity squared. We assume that the zero-albedo equilibrium temperature scales as $T_{\\rm eq} = T_{\\rm eff}(R_{\\star}/2a)^{1/2}$ (i.e., we assume that the heat redistribution factor $f$ is common to all planets, in the absence of more complete knowledge). We list $\\Theta$ and $T_{\\rm eq}$ for all transiting planets in Table~\\ref{tab:planetary}. \\cite{Hansen:07} pointed out a gap in the distribution of Safronov numbers, and defined Class~I planets as those with $\\Theta \\sim 0.07 \\pm 0.01$ and Class~II as $\\Theta \\sim 0.04 \\pm 0.01$. They tentatively proposed also that these two categories have other distinguishing characteristics, such as a difference in the average temperature of the host stars, or the orbital separations. Upon the discovery of HAT-P-5b \\citep{Bakos:07c} and HAT-P-6b \\citep{Noyes:07}, those authors pointed out that the distinction now seems less clear, as these two planets tend to fill the gap between Class~I and Class~II. The larger sample now available, and especially the more accurate and homogeneous set of properties presented here, offers the opportunity to revisit the issue. Figure~\\ref{fig:safronov} shows an updated version of the $\\Theta$--$T_{\\rm eq}$ diagram for transiting planets, which has some significant differences compared to the original version. Following \\cite{Hansen:07} we have excluded the massive planet HAT-P-2b, with a Safronov number so much larger than all the others ($\\Theta = 0.94$) that it would seem to be in a different class altogether, as well as GJ~436b, which is of much lower mass and a presumably different composition. \\begin{figure} \\vskip -0.35in \\epsscale{1.3} {\\hskip -0.2in\\plotone{f11.eps}} \\vskip -0.3in \\figcaption[]{Diagram of Safronov number ($\\Theta$) versus equilibrium temperature ($T_{\\rm eq}$) for transiting planets. Planet classes are labeled following \\cite{Hansen:07}. A tentative dividing line at $\\Theta = 0.05$ is indicated. \\label{fig:safronov}} \\end{figure} In our updated diagram the separation between Class~I and Class~II is still quite striking. The clustering of the Class~II objects around the value $\\Theta \\sim 0.04$ has tightened, if anything, and HAT-P-5b (\\#21) and HAT-P-6b (\\#22) do not encroach on the gap in Safronov numbers with Class~I. (The Class~II object with the lowest value of $\\Theta$ is OGLE-TR-111b [\\#6].) Thus, the dichotomy remains very suggestive. \\cite{Hansen:07} have argued that the principal distinction between the two classes is based on mass (planetary and/or stellar), and that planets of Class~II are, on average, less massive than those in Class~I, and orbit stars that are typically more massive. While we agree with the latter part of this statement regarding the \\emph{stellar} masses, we find that the distribution of \\emph{planetary} masses is indistinguishable between the two groups. We do confirm, however, that an updated diagram of $M_p$ versus $T_{\\rm eq}$ (see Figure~\\ref{fig:masstemp}) shows planets in Class~II to be systematically less massive \\emph{for the same equilibrium temperature} (level of irradiation), as was also found by \\cite{Hansen:07}, so in this sense their general claim that the distinction has something to do with mass seems to be supported by the observations. \\begin{figure} \\vskip -0.35in \\epsscale{1.3} {\\hskip -0.2in\\plotone{f12.eps}} \\vskip -0.3in \\figcaption[]{Planetary mass as a function of equilibrium temperature for transiting planets. Class~II planets ($\\Theta < 0.05$) are represented with squares. A tentative dividing line is indicated. \\label{fig:masstemp}} \\end{figure} An important characteristic that seems to be different in the two groups is the metallicity of the parent stars. This is illustrated in Figure~\\ref{fig:saffeh}. Parent stars of Class~II planets tend to be slightly more metal-rich. A Kolmogorov-Smirnov test of the two metallicity distributions, which appear to be centered around [Fe/H] $\\sim 0.0$ for Class~I and [Fe/H] $\\sim +0.2$ for Class~II, indicates only a 1.7\\% probability that they are drawn from the same parent population. There is perhaps a hint that the Safronov numbers for Class~I planets show a decreasing trend with metallicity in Figure~\\ref{fig:saffeh}, whereas the $\\Theta$ values for Class~II planets are independent of [Fe/H]. \\begin{figure} \\vskip -0.3in \\epsscale{1.3} {\\hskip -0.2in\\plotone{f13.eps}} \\vskip -0.3in \\figcaption[]{Safronov number as a function of stellar metallicity for transiting planets. Class~II planets ($\\Theta < 0.05$) are represented with squares.\\label{fig:saffeh}} \\end{figure} \\subsection{Heavy element content versus stellar metallicity} \\label{sec:cores} Transiting planet discoveries have challenged our theoretical understanding of these objects from the very beginning. The inflated radius of HD~209458b, argued to be due to an overlooked internal heat source in the planet \\citep[see, e.g.,][and references therein]{Fabrycky:07}, still poses a problem for modelers, and this first example is now joined by several other oversized planets indicating it is not an exception (see \\S\\,\\ref{sec:results}). On the other end of the scale, HD~149026b was the first transiting giant planet found to have a radius significantly \\emph{smaller} than predicted by standard theories \\citep{Sato:05}. The implication is that it must have a substantial fraction of heavy elements of perhaps $\\sim$70~M$_{\\earth}$ (2/3 of its total mass), which is often assumed to be in a core. More recently HAT-P-3b \\citep{Torres:07a} has also been found to be enhanced in heavy elements on the basis of its small size for the measured planet mass. In this case metals make up $\\sim$1/3 of the total mass. \\cite{Guillot:06} and also \\cite{Burrows:07} have found evidence that the heavy element content correlates with the metallicity of the parent star, a trend that was not anticipated by theory. These studies were based, respectively, on the 9 and 14 transiting planets known at the time. The larger sample now available warrants a second look at this possible correlation, for its potential importance for our understanding of planet formation. Here we have estimated the heavy element content $M_{\\rm Z}$ of each planet using the recent models by \\cite{Fortney:07}, which include the effects of irradiation from the central star. The mass in heavy elements was calculated from the measured mass of the planet, its radius, the orbital semimajor axis, and the age of the system as derived above from stellar evolution models. The uncertainty in these estimates is often very large, due mostly to errors in the measured values of $R_p$ and especially the age. In ten cases the models indicate no metal enhancement at all, and some of these are in fact ``inflated'' hot Jupiters. In a few other cases only an upper limit can be placed. For HAT-P-2b the \\cite{Fortney:07} models as published do not provide a large enough range of core masses. The results for $M_{\\rm Z}$ are listed in Table~\\ref{tab:planetary}. A comparison with values from \\cite{Guillot:06} and \\cite{Burrows:07} for the subset of planets in common indicates significant differences in some cases, either in $M_{\\rm Z}$ or its error. \\begin{figure} \\vskip -0.35in \\epsscale{1.3} {\\hskip -0.2in\\plotone{f14.eps}} \\vskip -0.3in \\figcaption[]{Heavy element content for transiting giant planets as a function of the metallicity of the host star. HAT-P-2 is excluded (see text).\\label{fig:cores}} \\end{figure} The mass in heavy elements is plotted against the stellar metallicity in Figure~\\ref{fig:cores}. The overall trend is similar to that pointed out by previous authors, in the sense that the upper envelope of the distribution appears to increase with [Fe/H]. It is natural to expect that a higher stellar metallicity implies a higher-metallicity protoplanetary disk. In the context of the core-accretion model of planet formation, one would also naturally expect a more metal-rich disk to lead to more metal-rich planets, and in this sense the observed trend is in accordance with the core-accretion theory." }, "0801/0801.3596_arXiv.txt": { "abstract": "This article investigates the predictions of an inflationary phase starting from a homogeneous and anisotropic universe of the Bianchi~$I$ type. After discussing the evolution of the background spacetime, focusing on the number of $e$-folds and the isotropization, we solve the perturbation equations and predict the power spectra of the curvature perturbations and gravity waves at the end of inflation. The main features of the early anisotropic phase is (1) a dependence of the spectra on the direction of the modes, (2) a coupling between curvature perturbations and gravity waves, and (3) the fact that the two gravity waves polarisations do not share the same spectrum on large scales. All these effects are significant only on large scales and die out on small scales where isotropy is recovered. They depend on a characteristic scale that can, but a priori must not, be tuned to some observable scale. To fix the initial conditions, we propose a procedure that generalises the one standardly used in inflation but that takes into account the fact that the WKB regime is violated at early times when the shear dominates. We stress that there exist modes that do not satisfy the WKB condition during the shear-dominated regime and for which the amplitude at the end of inflation depends on unknown initial conditions. On such scales, inflation loses its predictability. This study paves the way to the determination of the cosmological signature of a primordial shear, whatever the Bianchi~$I$ spacetime. It thus stresses the importance of the WKB regime to draw inflationary predictions and demonstrates that when the number of $e$-folds is large enough, the predictions converge toward those of inflation in a Friedmann-Lema\\^{\\i}tre spacetime but that they are less robust in the case of an inflationary era with a small number of $e$-folds. ", "introduction": "Inflation~\\cite{lindebook,pubook} (see Ref.~\\cite{linde2007} for a recent review of its status) is now one of the cornerstones of the standard cosmological model. In its simplest form, inflation has very definite predictions: the existence of adiabatic initial scalar perturbations and gravitational waves, both with Gaussian statistics and an almost scale invariant power spectrum~\\cite{ref:inf,mbf}. Various extensions, which in general involve more fields, allow e.g. for isocurvature perturbations~\\cite{iso}, non-Gaussianity~\\cite{ng}, and modulated fluctuations~\\cite{bku}. All these features let us hope that future data will shed some light on the details (and physics) of this primordial phase of the universe. Almost the entire literature on inflation assumes that the universe is homogeneous and isotropic while homogeneity and isotropy are what inflation is supposed to explain. Indeed, the dynamics of anisotropic inflationary universes has been widely discussed~\\cite{olive}. It was demonstrated that under a large variety of conditions, inflation occurs even if the spacetime is initially anisotropic~\\cite{infanisogen}, regardless of whether it is dominated by a pure cosmological constant or a slow-rolling scalar field. The isotropization of the universe was even generalized to Bianchi braneworld models~\\cite{braneinf} recently. It should be emphasized, however, that a deviation from isotropy~\\cite{infanisogen} or flatness~\\cite{inflK} may have a strong effect on the dynamics of inflation, and in particular on the number of $e$-folds.\\\\ The study of the perturbations during the isotropization phase has been overlooked, mainly because in the past decades one was mostly focused on large field inflationary models, which generically give very long inflationary phases. This was also backed up by the ideas of chaotic inflation and eternal inflation~\\cite{linde2007}. In such cases, it is thus an excellent approximation to describe the universe by a Friedmann-Lema\\^{\\i}tre (FL) spacetime when focusing on the modes observable today since they exited the Hubble radius approximatly in the last 60 $e$-folds. In this case, the origin of the density perturbations is understood as the amplification of vacuum quantum fluctuations of the inflaton. In particular, the degrees of freedom that should be quantized deep in the inflationary phase, and known as the Mukhanov-Sasaki variables~\\cite{MSvar}, were identified~\\cite{mbf}, which completely fixes the initial conditions and makes inflation a very predictive theory. In the context of string theory, constructing a string compactification whose low energy effective Lagrangian is able to produce inflation is challenging (see Ref.~\\cite{Allister}). In particular, it has proved to be difficult to build large field models~\\cite{bauman} (in which the inflaton moves over large distances compared to the Planck scale in field space). This has led to the idea that, in this framework, an inflationary phase with a small number of $e$-folds is favored (see however Ref.~\\cite{roulette}). If so, the predictions of inflation are expected to be sensitive to the initial conditions, and in particular the classical inhomogeneities are not expected to be exponentially suppressed, which makes the search for large scale deviations from homogeneity and isotropy much more motivated, as well as the possibility that the inflaton has not reached the inflationary attractor. More important, it is far from obvious (as we shall discuss in details later) that all observable modes can be assumed to be in their Bunch-Davies vacuum initially, and more puzzling, that for a given mode modulus the possibility of setting the initial conditions will depend on its direction. If so, then the initial conditions for these modes would have to be set in the stringy phase, an open issue at the time. The theory of cosmological perturbations in a Bianchi universe was roughed out in Ref.~\\cite{TomitaDen,Dunsby,pitrou07} (see also Ref.~\\cite{NohHwang} and Ref.~\\cite{Abbotetal} for the case of higher-dimensional Kaluza-Klein models and Ref.~\\cite{Qaniso} for the quantization of test fields and particle production in an anisotropic spacetime). Recently, we performed a full analysis of the cosmological perturbations in an arbitrary Bianchi~$I$ universe~\\cite{ppu1}. It was soon followed by an analysis~\\cite{GCP} that focused on Bianchi~$I$ universe with a planar symmetry. As we shall see in this work, the case of a planar symmetric spacetime is not generic (both for the dynamics of the background and the evolution of the perturbations).\\\\ From a more observationally oriented perspective, the primordial anisotropy can imprint a preferred direction in the primordial power spectra. This could be related to the possible large scale statistical anomalies~\\cite{lowQ} of the cosmic microwave background (CMB) anisotropies. Many possible explanations have been proposed, including foregrounds~\\cite{prunet}, non-trivial spatial topology~\\cite{topologie} (which implies a violation of global isotropy~\\cite{modetopo}), the breakdown of local isotropy due to multiple scalar fields~\\cite{picon}, the presence of spinors~\\cite{spinor} or dynamical vectors \\cite{dulaney}, the effect of the spatial gradient of the inflaton~\\cite{donoghue} or a late time violation of isotropy~\\cite{Jaffe2005}. In the case of universes with planar symmetry, the signatures on the CMB were derived by many authors~\\cite{GCP,CMB_bianchi,acker}. In this case, a large primordial shear is indeed necessary, which is not in contradiction with the constraints obtained from the CMB~\\cite{LimitShear} or from the big-bang nucleosynthesis~\\cite{bbn}. This brings a secondary motivation to our analysis: can the CMB anisotropy be related to an anisotropic primordial phase or, on the other hand, can it constrain the primordial shear and what are the exact predictions of a primordial anisotropic phase? \\\\ This is however not our primary motivation. In the first place, we are interested in understanding the genericity of the predictions of inflation, and in particular with respect to the symmetries of the background spacetime. As we shall see, the simple extension considered in this article drives a lot of questions, concerning both the initial conditions in inflation and, more generally, quantum field theory in curved spacetime. In this article, we build on our previous work~\\cite{ppu1} to investigate the dynamics and predictions of one field inflation starting from a generic Bianchi~$I$ universe. After discussing the dynamics of the background in \\S~\\ref{sec1}, we focus on the evolution of the perturbations in \\S~\\ref{sec2}. In particular, we will need to understand the procedure for the quantization during inflation. As we shall see in~\\S~\\ref{sec2.2}, this procedure deviates from the one standardly used in a Friedmann-Lema\\^{\\i}tre spacetime, and even in a planar symmetric spacetime as discussed in Ref.~\\cite{GCP}. The main reason for such an extension lies in the fact that there always exist modes that were not in a WKB regime during the shear-dominated inflationary era. It follows that, while our procedure leads to similar predictions to the standard one on small scales, it appears that there is a lack of predictibility on large scales. Indeed, we do not want to push our description beyond the Planck or string scale where extensions of general relativity have to be considered. This may give a description of both the early phase of the inflationary era and also a procedure to fix the initial conditions without any ambiguity. In \\S~\\ref{sec3}, we explicitly compute the primordial spectra, for both scalar modes and gravity waves. We will describe in details the effect of the anisotropy and show how the isotropic predictions are recovered on small scales. ", "conclusions": "\\label{sec4} In this article, we have worked out the predictions of an anisotropic inflationary era for a generic Bianchi~$I$ spacetime. We have discussed the isotropization both at the background level and at the linear order in perturbation theory, both for scalar modes and gravity waves. Generically these spacetimes always enjoy a bouncing direction, apart from the particular case $\\alpha=\\pi/2$ considered in Ref.~\\cite{GCP}. (Note that the predictions for this case are in fact singular among the predictions since they do not converge uniformly when $\\alpha\\rightarrow\\pi/2$). Since at early time, the modes are not in a WKB regime, we had to extend the standard procedure to fix the initial conditions (see \\S~\\ref{sec2.2.c}). We showed that the modes larger than $\\kref$ always enter a WKB regime before they become super-Hubble during inflation. As we discussed, our procedure reproduces the standard one on small scales. In the particular case where we tune the initial shear so that these initial conditions can be set unambiguously while still having an imprint of the CMB anisotropy, we presented the imprint of the primordial anisotropy in the power spectra of the gravity waves and curvature perturbation at the end of inflation. Two examples were studied but we can provide predictions for any Bianchi~$I$ universe. Note that these predictions were drawn by assuming that the slow-roll attractors were reached before the time the modes of observational relevance had exited the horizon. In the case of inflation with small number of $e$-folds before that time, it is not clear that this is a realistic hypothesis. If so, one would have to make the predictions in terms of trajectories, that is predictions that will depend on the initial conditions of the scalar field. This is not specific to Bianchi universes but to all models in which the inflationary period is short~\\cite{levprivate}.\\\\ Concerning the initial conditions, two problems arise (see discussion in \\S~\\ref{sec2.2.e}). First, there exists an early shear-dominated phase where the WKB approximation is violated. This forbids us to set the initial conditions as in a Friedmann-Lema\\^{\\i}tre universe. As we showed, the sub-Hubble modes at the onset of the accelerating phase can be quantized because quantum fluctuations act at all times and can always source oscillatory solutions, when they exist. On the other hand, there always exist non-oscillating modes. These modes are expected not to be much smaller than the Hubble radius today (since there is no trivial imprint of the anisotropy on the CMB). This implies that there exists a cut-off scale above which we cannot predict the spectra from first principles, at least in the theoretical set-up we are considering. Consequently, we conclude that above this scale we can only measure the power spectra or postulate their functional form, as was actually done to set the initial conditions before the invention of inflation. We have shown that even if unknown pre-WKB initial conditions can grow, this growth is at most of order unity. Therefore we can safely assume that the perturbations at the end of inflation reflect only those modes that have been seeded during the WKB regime. Second, for these modes, one cannot assume that they are independent. It implies that we have to consider 3 interacting fields, which also complicates the quantization procedure. Such an issue was addressed perturbatively in the interaction picture in the case of self-interacting field in order to estimate the non-Gaussianity~\\cite{interacting} but no general formalism has been designed when no interaction-free regime can be exhibited.\\\\ In our analysis we have assumed the validity of general relativity up to the singularity. Indeed, we do not take this model for more than what it actually is, e.g. we cannot extrapolate it beyond the Planck or the string scale. There, more degrees of freedom have to be included and can change the dynamics. This could introduce an early chaotic phase~\\cite{billiard} since Bianchi~$I$ models coupled to $p$-form fields have never ending oscillatory behaviour exhibited by generic string theory. Examples of an early dynamics have also been given in terms of Kalb-Ramond axion fields~\\cite{kaloper}, non-commutative geometry~\\cite{ncg} and recently loop quantum gravity~\\cite{lcq}. Any of these developments may give a description of both the early phase and a procedure to fix the initial conditions.\\\\ Coming back to inflation, our study demonstrates to which extent its predictions are sensitive to initial (classical) large scale anisotropies and that in the presence of a non-vanishing shear, it is impossible to define a Bunch-Davies vacuum in the standard way (see also Ref.~\\cite{picon2} for a discussion of this issue in the standard picture and the arbitrariness on the choice of the initial state and its influence of the prediction of inflation). Indeed, if we tune the initial conditions such that the number of $e$-folds is large, none of the problems we address here will affect the observable modes, simply because only modes such that $k/\\kref\\gg1$ are observable and we have shown that in this limit we recover the standard inflationary predictions. If this is the case, one would have no observational imprint of the primordial shear. But our analysis and conclusions may be of some relevance for inflationary model building in the framework of string theory if the feeling that no large field model (and thus no large number of $e$-folds) can be constructed persists, an issue far beyond the scope of this article. Our analysis also shows the importance (and peculiarity) of the Friedmann-Lema\\^{\\i}tre background in our theoretical predictions and on the quantization procedure, and it gives as well an explicit construction of the difficulties encountered when these symmetries do not exist. We showed that if the number of $e$-folds is large the inflationary predictions converge toward the isotropic predictions, hence demonstrating that they are robust in that regime. In the case of a small number of $e$-folds, they are very sensitive to the initial shear but also the theoretical construction is less under control. If the indication of the breakdown of statistical isotropy from the CMB were to be confirmed and related to such an early anisotropic phase, then a new coincidence will appear in the cosmological models since one would need to understand why the characteristic scale is of the order of the Hubble radius today, $\\kref\\sim k_0$. From a more pragmatic attitude, the present work allows us to draw the CMB signatures of such an anisotropic early phase, for any Bianchi~$I$ universe. We stress that the general form of the initial power spectra are more general than those heuristically considered in the analysis performed in Refs.~\\cite{acker,kam} and go beyond the one derived for the singular case $\\alpha=\\pi/2$ studied in Ref.~\\cite{GCP}. The existence of correlation between gravity waves and curvature perturbation, as well as the fact that the two gravity wave polarisations do not share the same spectrum, may also lead to specific signatures to be investigated." }, "0801/0801.3075_arXiv.txt": { "abstract": "Circumnuclear star forming regions, also called hotspots, are often found in the inner regions of some spiral galaxies where intense processes of star formation are taking place. In the UV, massive stars dominate the observed circumnuclear emission even in the presence of an active nucleus, contributing between 30 and 50 \\% to the H$\\beta$ total emission of the nuclear zone. Spectrophotometric data of moderate resolution ( 3000 $<$ R $<$ 11000) are presented from which the physical properties of the ionized gas: electron density, oxygen abundances, ionization structure etc. have been derived. ", "introduction": "The inner ($\\sim$ 1Kpc) zones of some spiral galaxies show high star formation rates frequently arranged in a ring or pseudo-ring pattern around their nuclei. In general, Circumnuclear Star Forming Regions (CNSFR) -- also referred to as ``hotspots'' -- and luminous and large disk HII regions are very much alike, but the first look more compact and show higher peak surface brightness (Kennicutt et al. 1989). In many cases they contribute substantially to the emission of the entire nuclear region. Their large H$\\alpha$ luminosities, typically higher than 10$^{39}$ erg s$^{-1}$, point to relatively massive star clusters as their ionisation source. % These regions then constitute excellent places to study how star formation proceeds in high metallicity, high density circumnuclear environments. The importance of an accurate determination of the abundances of high metallicity HII regions cannot be overestimated since they constitute most of the HII regions in early spiral galaxies (Sa to Sbc) and the inner regions of most late type ones (Sc to Sd) without which our description of the metallicity distribution in galaxies cannot be complete. The question of how high is the highest oxygen abundance in the gaseous phase of galaxies is still standing and extrapolation of known radial abundance gradients would point to CNSFR as the most probable sites for these high metallicities. Accurate measurements of elemental abundances of high metallicity regions are therefore crucial to obtain reliable calibrations of empirical abundance estimators, widely used but poorly constraint, whose choice can severely bias results obtained for quantities of the highest relevance for the study of galactic evolution like the luminosity-metallicity (L-Z) relation for galaxies. CNSFR are also ideal cases for studying the behaviour of abundance estimators in the high metallicity regime. ", "conclusions": "Electron densities for each observed region have been derived from the [SII] $\\lambda\\lambda$ 6717, 6731 \\AA\\ line ratio, following standard methods. They were found to be, in all cases, $\\le$ 600 cm$^{-3}$, higher than those usually derived in disk HII regions, but still below the critical value for collisional de-excitation. The low excitation of the regions, as evidenced by the weakness of the [OIII] $\\lambda$ 5007 \\AA\\ line (see the blue spectrum in Figure 1, precludes the detection and measurement of the auroral [OIII] $\\lambda$ 4363 \\AA\\ necessary for the derivation of the electron temperature. We have therefore used a semi-empirical procedure for the derivation of abundances. Firstly we have produced a calibration of the [SIII] temperature as a function of the abundance parameter SO$_{23}$ defined as: \\[ SO_{23} = \\frac{I([OII]\\lambda 3727,29)+I([OIII]\\lambda 4959,5007)}{I([SII]\\lambda 6716,31)+I([SIII]\\lambda 9069,9532)} \\] This parameter is similar to the S$_3$O$_3$ proposed by Stasi\\'nska (2006) but is, at first order, independent from geometrical (ionization parameter) effects. To perform this calibration we have compiled all the data so far available with sulfur emission line detections of both the auroral and nebular lines at $\\lambda$ 6312 \\AA\\ and $\\lambda\\lambda$ 9069,9532 \\AA\\ respectively (see D\\'\\i az et al. 2007). The calibration is shown in Figure 2 together with the quadratic fit to the high metallicity HII region data: \\[ t_e([SIII]) = 0.596 - 0.283 log SO_{23} + 0.199 (log SO_{23})^2 \\] For one of the observed regions, R1+R2, the [SIII] $\\lambda$ 6312 \\AA\\ line has been detected and measured yielding a value T$_e$([SIII])= 8400$^{+ 4650}_{-1250}$K, represented as a solid black circle in Figure 2 (left). We have used this calibration to derive t$_{e}$([SIII]) for our observed CNSFR. In all cases the values of log$O_{23}$ are inside the range used in performing the calibration, thus requiring no extrapolation of the fit.These temperatures, in turn, have been used to derive the S$^+$/H$^+$ and S$^{++}$/H$^+$ ionic ratios. Once the sulphur ionic abundances have been derived, we have estimated the corresponding oxygen abundances, assuming that sulphur and oxygen temperatures follow the relation given by Garnett (1992) and t$_e$ ([OII]) $\\simeq$ t$_e$([SIII]). Finally, we have derived the N$^{+}$/O$^{+}$ ratio assuming that t$_e$([OII]) $\\simeq$ t$_e$([NII]) $\\simeq$ t$_e$([SIII]). \\begin{figure}[!h] \\plottwo{TeSIII_cal_new.eps}{eta-prima-plot.eps} \\caption{{\\itshape Left:\\/}Empirical calibration of the [SIII] electron temperature as a function of the abundance parameter SO$_{23}$. The solid line represents a quadratic fit to the high metallicity HII region data. {\\itshape Right:\\/} The [OII]/[OIII] vs [SII]/[SIII] logarithmic ratios for different ionised regions. References for the data in the two figures can be found in D\\'\\i az et al. (2007)} \\end{figure} The observed CNSFR being of high metallicity show however marked differences with respect to high metallicity disk HII regions. Even though their derived oxygen and sulphur abundances are similar, they show values of the O$_{23}$ and the N2 parameters whose distributions are shifted to lower and higher values respectively with respect to the high metallicity disk sample. Hence, if pure empirical methods were used to estimate the oxygen abundances for these regions, higher values would in principle be obtained. This would seem to be in agreement with the fact that CNSFR, when compared to the disk high metallicity regions, show the highest [NII]/[OII] ratios. CNSFR also show lower ionization parameters than their disk counterparts, as derived from the [SII]/[SIII] ratio. Their ionization structure also seems to be different with CNSFR showing radiation field properties more similar to HII galaxies than to disk high metallicity HII regions. This can be seen in Figure 2 (right), a diagram of the emission line ratios [OII]/[OIII] vs [SII]/[SIII], that works as a diagnostics for the nature and temperature of the radiation field. In this plot, diagonal lines of slope unity show the locus of ionized regions with constant ionization temperature. CNSFRs are seen to segregate from disk HII regions. The former cluster around a value of T$_{ion}\\sim$ 40,000 K, while the latter cluster around T$_{ion}\\sim$ 35,000 K. Also shown are the data corresponding to HII galaxies. Indeed, CNSFRs seem to share more the locus of HII galaxies than that of disk HII regions. One possible concern about these CNSFR is that, given their proximity to the galactic nuclei, they could be affected by hard radiation coming from a low luminosity AGN. Alternatively, the spectra of these regions harbouring massive clusters of young stars might be affected by the presence of shocked gas. Diagnostic diagramas of the kind presented by Baldwin, Phillips, \\& Terlevich (1981) can be used to investigate the possible contribution by either a hidden AGN or the presence of shocks to the emission line spectra of the observed CNSFR. When plotted on one of these diagnostic diagramas, log ([NII]/H$\\alpha$) vs log ([OIII])/H$\\beta$, some of our CNSFR are found close to the transition zone between HII region and LINER spectra but only one region in NGC~3504 may show a hint of a slight contamination by shocks." }, "0801/0801.3305_arXiv.txt": { "abstract": "We investigate the accretion of angular momentum onto a protoplanet system using three-dimensional hydrodynamical simulations. We consider a local region around a protoplanet in a protoplanetary disk with sufficient spatial resolution. We describe the structure of the gas flow onto and around the protoplanet in detail. We find that the gas flows onto the protoplanet system in the vertical direction crossing the shock front near the Hill radius of the protoplanet, which is qualitatively different from the picture established by two-dimensional simulations. The specific angular momentum of the gas accreted by the protoplanet system increases with the protoplanet mass. At Jovian orbit, when the protoplanet mass $M_{\\rm p}$ is $M_{\\rm p}\\lesssim 1\\mj$, where $\\mj$ is Jovian mass, the specific angular momentum increases as $j\\propto M_{\\rm p}$. On the other hand, it increases as $j\\propto M_{\\rm p}^{2/3}$ when the protoplanet mass is $M_{\\rm p}\\gtrsim 1\\mj$. The stronger dependence of the specific angular momentum on the protoplanet mass for $M_{\\rm p}\\lesssim 1\\mj$ is due to thermal pressure of the gas. The estimated total angular momentum of a system of a gas giant planet and a circumplanetary disk is two-orders of magnitude larger than those of the present gas giant planets in the solar system. A large fraction of the total angular momentum contributes to the formation of the circumplanetary disk. We also discuss the satellite formation from the circumplanetary disk. ", "introduction": "Until now, more than 200 extrasolar planets (or exoplanets) have been detected mainly by measuring the radial motion of their parent star along the line of sight. Almost all exoplanets observed by this method are giant planets, like Jupiter and Saturn in our solar system, because massive planets are preferentially observed. Although these planets are supposed to be formed in the disk surrounding the central star (i.e., circumstellar disk or protoplanetary disk), their formation process has not been fully understood yet. In the core accretion scenario \\citep{hayashi85}, a solid core or protoplanet with $\\simeq 10\\me$, where $\\me$ is the Earth mass, captures a massive gas envelope from the protoplanetary disk by self-gravity to become a gas giant planet. The evolution of the gaseous protoplanet has been studied with the approximation of spherical symmetry including radiative transfer \\citep[e.g.,][]{mizuno80,bodenheimer86,pollack96,ikoma00}. \\citet{ikoma00} showed that rapid gas accretion is triggered when the solid core mass exceeds $\\simeq 5-20\\me$, and the protoplanet quickly increases their mass by gas accretion. However, the angular momentum of the accreting gas was ignored in these studies, because they assumed spherical symmetry. Since the gas accretes onto the solid core with a certain amount of the angular momentum, a circumplanetary disk forms around the protoplanet, analogously to the formation of a protoplanetary disk around a protostar. The difference in the disk formation between the protostar and protoplanet is the region from which the central object acquires the angular momentum. The protostar acquires the angular momentum from a parent cloud, while the protoplanet acquires it from the shearing motion in the protoplanetary (circumstellar) disk. In addition, the gravitational sphere of the protostar spreads almost infinitely, while the gravitational sphere (i.e., the Hill sphere) of the protoplanet is limited in the region around the protoplanet because the gravity of the central star exceeds that of the protoplanet outside the Hill sphere. The numerical simulations are useful to investigate the gas accretion onto a protoplanet and its circumplanetary disk (hereafter we just call them as a protoplanet system). \\citet{korycansky91} studied giant planet formation using one-dimensional quasi-spherical approximation with angular momentum transfer. They showed that as the protoplanet contracts, outer layers of the envelope containing sufficient specific angular momentum remain in bound orbit, and form a circumplanetary disk. However, owing to the spherical symmetry, the accretion flow from the protoplanetary disk to the protoplanet could not be investigated in their study. To study the accretion flow onto a protoplanet and the acquisition process of the angular momentum in detail, a multidimensional simulation is necessary. \\citet{sekiya87} investigated the gas flow around a protoplanet with relatively low resolution, and found that the spin rotation vector of the protoplanet becomes parallel to the orbital rotation vector. Recently, the flow pattern was carefully investigated by many authors \\citep{miyoshi99, lubow99, kley01, dangelo02, dangelo03}. However, since the main purpose of these studies was to clarify the planet migration process in a large scale (i.e., outside the Hill radius), they did not investigate the region near the protoplanet (i.e., inside the Hill radius) with sufficient resolution. Thus, in their simulations, we cannot study the gas stream inside the Hill radius. To investigate the accretion flow onto the protoplanet system, we need to cover a large spatial scale from the region far from the Hill sphere to that in the proximity to the protoplanet. For Jupiter, since the Hill radius is $\\rh = 744\\,\\rp$, where $\\rp$ is Jovian radius, we have to resolve at least $\\sim$1000 times different scales. To cover a large dynamical range in scales, a few authors used the nested-grid method. \\citet{dangelo02,dangelo03} investigated the relation between the spiral patterns within the Hill radius and migration rate using three-dimensional nested-grid code. Although they resolved the region inside the Hill radius, they did not investigate the structure in the proximity to the protoplanet because they adopted a sink cell at $0.1r_{\\rm H}$ that corresponds to $\\sim70\\,\\rp$ at Jovian orbit (5.2 AU). Thus, we cannot observe a circumplanetary disk at $r\\ll\\rh$ in their calculation. \\citet{tanigawa02a} also investigated the gas flow around a protoplanet using two-dimensional nested-grid code. They resolved the region from $12\\,\\rh$ to $0.005\\,\\rh$. They found that a circumplanetary disk with $100\\me$ is formed around the protoplanet. \\citet{tanigawa02b} and \\citet{machida06c} investigated the evolution of the protoplanet system using three-dimensional nested-grid code. They found that the gas flow pattern in three dimensions is qualitatively different from that in two dimensions: the gas is flowing into the protoplanet system only in the vertical direction in three-dimensional simulations. In the present study, we calculated the evolution of the protoplanet system using three-dimensional nested-grid code. We found that after the flow around the protoplanet reaches a steady state, the angular momentum accreting onto the protoplanet system is well converged regardless of both the cell width and the size of the sink cell region, while the mass accretion rate is not well converged. Although we calculated the evolution of the protoplanet system with spatial resolution much higher than previous studies, we still need further higher spatial resolution to determine the mass accretion rate onto the protoplanet. Thus, in this paper, we focus on the gas flow onto and around the protoplanet system and the accretion process of the angular momentum, and do not deal with the mass accretion rate (we plan to investigate the mass accretion rate with higher spatial resolution using a higher-performance computer in a subsequent paper). Note that the specific angular momentum accreting onto the protoplanet system with fixed protoplanet mass does not strongly depend on the mass accretion rate as described in the following sections. The structure of the paper is as follows. The frameworks of our models are given in \\S 2 and the numerical method is described in \\S3. The numerical results are presented in \\S 4. \\S 5 is devoted for discussions. We summarize our conclusions in \\S6. ", "conclusions": "\\subsection{Effect of Thermal Pressure} \\label{sec:thermal} In this paper, we investigated the evolution of the protoplanet system under the isothermal approximation, which might not be problematic in the circumplanetary disk, but may not be valid in the very proximity to the protoplanet. To investigate the thermal effect around the protoplanet, we adopted the sink cell in some models. As shown in \\S\\ref{sec:sink}, the cell width in the fiducial model having $l_{\\rm max}=8$ is about 4 Jovian radius. Thus, gas inside $r < 4 \\rj$ has the same thermal energy. If the protoplanet has almost the same size as the present gas giant planet, the thermal energy around the protoplanet may be overestimated in model without the sink cell, because an actual protoplanet is embedded in the small part of the innermost cell. On the other hand, when we adopt the sink cell, the thermal energy around the protoplanet is underestimated, because the thermal energy is artificially removed from the sink cell. However, as shown in \\S\\ref{sec:sink}, when the sink radius is much smaller than the Hill radius ($r_{\\rm sink} \\ll 1/100 \\rh$), there are little differences in the angular momentum acquired by the protoplanet system between models with and without the sink cell. This is because a large part of the angular momentum of the protoplanet system is distributed around the Hill radius ($0.5\\,\\rh\\lesssim r \\lesssim \\rh$) as shown in \\S\\ref{sec:res-ang}. The Jovian radius is much smaller than the Hill radius ($\\rh = 744\\, \\rj$). In addition, the angular momentum flowing into the protoplanet system is determined by the shearing motion in the region of $r \\simeq \\rh$ (see \\ref{sec:dis-evo}). Therefore, under the assumption that the protoplanet is smaller than the innermost cell width, it is expected that the thermal effect from the protoplanet is sufficiently small for the acquisition process of the angular momentum when we are using sufficient smaller cells than the Hill radius. \\citet{mizuno80}, \\citet{bodenheimer86}, and \\citet{ikoma00} suggested that gaseous protoplanet has a large envelope with high temperature. For example, \\citet{mizuno80} showed that, for Jovian case, gas distributed in the range of $r\\gtrsim 2\\times 10^{11}$\\,cm behaves isothermally, while that distributed in the range of $r \\lesssim 2\\times 10^{11}$\\,cm behaves adiabatically in their spherically symmetric calculations (see Fig.4 of \\citealt{mizuno80}). Thus, gas in the range of $r \\lesssim 2\\times 10^{11}$\\,cm ($\\sim30\\,\\rj$) has higher temperature than the ambient medium. However, it is expected that the gaseous envelope hardly affect the angular momentum of the protoplanet system, because the size of the envelope ($\\sim30\\,\\rj$) is much smaller than the Hill radius ($\\rh = 744\\, \\rj$). A large part of the angular momentum is distributed in the range of $0.5\\rh \\lesssim r \\lesssim \\rh$, as shown in \\S\\ref{sec:res-ang}. On the other hand, the size of the circumplanetary disk may be affected by the thermal envelope. Thus, when we investigate the formation of the circumplanetary disk, we have to include the realistic thermal evolution around the protoplanet. However, it is difficult to study the thermal evolution around the protoplanet, because we need to solve the radiation hydrodynamics in three dimensions. In a subsequent paper, we will investigate the effect of the thermal envelope under simple assumptions. \\subsection{Angular Momentum for Jupiter and Saturn} \\label{sec:jupiter-saturn} In \\S\\ref{sec:dis-evo}, we fixed the physical quantities as those at $\\ro=5.2$\\,AU. However, in our calculation, since we use the dimensionless quantities, we can rescale those at any orbits $\\ro$. Under the standard model \\citep{hayashi85}, we can generalize equation~(\\ref{eq:jlm}) as \\begin{equation} j_{\\rm lm} = 7.8 \\times 10^{15} \\left(\\dfrac{M_{\\rm p}}{M_{\\rm J}} \\right) \\left(\\dfrac{\\ro}{1\\,{\\rm AU}} \\right)^{7/4} \\jcm = 2.5 \\times 10^{13} \\left( \\dfrac{M_{\\rm p}}{\\me} \\right) \\left( \\dfrac{\\ro}{1\\, {\\rm AU}} \\right)^{7/4} \\jcm. \\label{eq:fit3} \\end{equation} We assume that the protoplanet system acquires the gas with the average specific angular momentum of equation~(\\ref{eq:fit3}) when the protoplanet has a mass of $M_{\\rm p}<1\\mj$. Thus, to estimate the angular momentum of the protoplanet system, we need to integrate equation~(\\ref{eq:fit3}) by the mass until the present values of gas giant planets. For example, the angular momentum in our model for Jupiter is \\begin{equation} J_{\\rm J} = \\int^{\\mj} j_{\\rm lm} (5.2\\,{\\rm AU})\\, dM = 1.3 \\times 10^{47} {\\rm g\\jcm}, \\end{equation} which is about 30 times larger than that of present Jupiter ($4.14\\times 10^{45}$g$\\jcm$). On the other hand, the angular momentum in our model for Saturn is \\begin{equation} J_{\\rm S} = \\int^{M_{\\rm s}} j_{\\rm lm} (9.6\\,{\\rm AU}) \\, dM = 3.6\\times 10^{46} \\ \\ {\\rm g\\,\\jcm}, \\end{equation} which is 50 times larger than that of present Saturn ($7.2 \\times 10^{44}$g$\\jcm$), where $M_{\\rm s} = 95.16\\me$ is the Saturnian mass. Note that the orbital angular momenta of the {\\em present} Jovian and Saturnian satellites can be ignored in the protoplanet system because they are considerably smaller than the Jovian and Saturnian spin angular momenta. The above estimation corresponds to the angular momentum of proto planet-disk system, which includes the central planet and protoplanetary disk. We showed that the angular momentum of the protoplanet system is $30-50$ times larger than present spin of the gaseous planet. The angular momenta flowing into the Hill sphere are distributed into the spin of the protoplanet and orbital motion of the circumplanetary disk. It is expected that a large fraction of the total angular momentum contributes to the formation of the circumplanetary disk and the residual contributes to the spin of the planet. Although the fraction of the angular momentum of the circumplanetary disk to the spin of protoplanet is not correctly estimated in our calculation, a part of the angular momentum is certainly distributed into the disk. Thus, the angular momentum transfer and dissipation mechanism for the circumplanetary disk are necessary to follow further evolution of the protoplanet system. \\citet{takata96} proposed the despin mechanism of the protoplanet by planetary dipole magnetic field, in which an initially rapidly rotating protoplanet can be spun down to the present value by the magnetic interaction between the protoplanet and circumplanetary disk. \\subsection{Disk formation and Implication for Satellite Formation} \\label{sec:dis-disk} When the circumplanetary disk is formed around the protoplanet, it is possible to form satellites in the disk. While there are many scenarios for the satellite formation \\citep[e.g.,][]{stevenson86}, regular satellites around gas giant planets are supposed to be formed in the gaseous disk as for the planet formation in the protoplanetary disk \\citep[e.g.,][]{korycansky91, canup02}. Observations showed that the regular satellites around Jupiter and Saturn are distributed only in the close vicinity of the planet ($r\\lesssim 50\\,\\rp$), and are on prograde orbits near the equatorial plane. Thus, it is expected that these regular satellites formed in the circumplanetary disk. We discussed the angular momentum of the protoplanet system in \\S\\ref{sec:dis-evo}. As the angular momentum flowing into the Hill sphere is brought into both the protoplanet and circumplanetary disk, we cannot estimate the fraction of angular momentum for the circumplanetary disk. When we assume that the centrifugal force is balanced with the gravity of the protoplanet, we can derive the centrifugal radius $r_{\\rm cf}$ as \\begin{equation} r_{\\rm cf} = \\dfrac{j^2}{GM_{\\rm p}}. \\label{eq:cent} \\end{equation} To quantify the disk size, we adopt the specific angular momentum in equation~(\\ref{eq:cent}) as those in equation~(\\ref{eq:fit3}). The centrifugal radius for the proto-Jovian disk with $M_{\\rm p} = 1 \\mj$ and $\\ro =5.2$\\,AU is \\begin{equation} r_{\\rm cf, J} = 1.5 \\times10^{11} \\ \\ {\\rm cm}, \\label{eq:r-est} \\end{equation} which is 22 times as large as Jovian radius ($r_{\\rm Jup} = 7.1\\times 10^9\\cm$). The Galilean satellites, which are regular satellites, are distributed in the range of $6\\,r_{\\rm Jup}\\lesssim r \\lesssim 27\\,r_{\\rm Jup}$, which is consistent with the centrifugal radius derived from our calculation. Note that the disk size is expected to be larger than equation~(\\ref{eq:r-est}) owing to the thermal effect around the protoplanet. In the same way, we estimate the centrifugal radius of proto-Saturnian disk as \\begin{equation} r_{\\rm cf, S} = 4.1\\times10^{11} \\ \\ {\\rm cm}, \\end{equation} which is 68 times as large as Saturnian radius ($r_{\\rm Sat}=6.0\\times10^9$\\,cm). The Saturnian representative regular satellites (Mimas, Enceladus, Tethys, Dione, Rhea, Titan, Hyperion, and Iapetus) are distributed in the range of $3\\, r_{\\rm Sat} \\lesssim r \\lesssim 60\\, r_{\\rm Sat}$, which is also corresponding to the centrifugal radius of our result. Figure~\\ref{fig:13} shows the structure around the protoplanet at the end of the calculation ($\\deft \\simeq 20$) for model M1, in which the protoplanet has mass of $1\\mj$. Each right panel is 4 times magnification of each left panel. Figures~\\ref{fig:13}{\\it a}, {\\it b}, and {\\it c} show the shock fronts outside the Hill radius, and a thick disk between the shock front and the Hill sphere. In the proximity to the protoplanet, a concave structure is seen in Figures~\\ref{fig:13}{\\it e} and {\\it f}. The contours in these figures rapidly drop around the rotation axis (i.e., $z$-axis). Thus, even inside the Hill sphere, except for the central region, the disk has a thick torus-like structure. Figure~\\ref{fig:13}{\\it g}, {\\it h}, and {\\it i} show the structure in the very proximity to the protoplanet. On the midplane, the elliptical structure is seen in a large scale (Fig.~\\ref{fig:13}{\\it d}), while an almost round shape is seen in the proximity to the protoplanet (Fig.~\\ref{fig:13}{\\it g}) because the force acting on gas in this region is dominated by the gravity of planet. The contours in Figures~\\ref{fig:13}{\\it h} and {\\it i} show a very thin disk in the region of $\\tl{r} \\lesssim 0.1$ (or $r \\lesssim 50\\,\\rj$). Thus, the regular satellites might be formed in these circumplanetary disks. However, to study the satellite formation in more detail, we need more realistic calculations." }, "0801/0801.3133_arXiv.txt": { "abstract": "{I review briefly different aspects of the MOND paradigm, with emphasis on phenomenology, epitomized here by many MOND laws of galactic motion--analogous to Kepler's laws of planetary motion. I then comment on the possible roots of MOND in cosmology, possibly the deepest and most far reaching aspect of MOND. This is followed by a succinct account of existing underlying theories. I also reflect on the implications of MOND's successes for the dark matter (DM) paradigm: MOND predictions imply that baryons alone accurately determine the full field of each and every individual galactic object. This conflicts with the expectations in the DM paradigm because of the haphazard formation and evolution of galactic objects and the very different influences that baryons and DM are subject to during the evolution, as evidenced, e.g., by the very small baryon-to-DM fraction in galaxies (compared with the cosmic value). All this should disabuse DM advocates of the thought that DM will someday be able to reproduce MOND: it is inconceivable that the modicum of baryons left over in galaxies can be made to determine everything if a much heavier DM component is present.} ", "introduction": "\\label{section1} MOND is an alternative paradigm to Newtonian dynamics, whose original motivation was to explain the mass discrepancies in galactic systems without invoking dark matter (DM) (Milgrom 1983a). It constitutes a modification of dynamics in the limit of low accelerations that rests on the following basic assumptions: (i) There appears in physics a new constant, $\\az$, with the dimensions of acceleration. (ii) Taking the formal limit $\\az\\rar 0$ in all the equations of physics restores the equations of classical (pre-MOND) dynamics. (iii) For purely gravitational systems, the opposite, deep-MOND limit, $\\az\\rar \\infty$, gives limiting equations of motion that can be written in a form where the constants $\\az$ and $G$, and all masses in the problem, $m_i$, appear only in the product $m_iG\\az=m_i/\\mz$, where $\\mz\\equiv(G\\az)^{-1}$ (Milgrom 2005)\\footnote{By this I mean that starting with equations that involve (including in derivatives) $\\vr$, $t$, $G$, $\\az$, $m_i$, and gravitational field degrees of freedom, we can rewrite them, possibly by redefining the gravitational field, so that only $\\vr$, $t$, $m_i/\\mz$, and the gravitational field appear.}. This last fiat reproduces the desired MOND phenomenology for purely gravitational systems. A MOND theory is one that incorporates the above tenets in the nonrelativistic regime. \\par Since all our knowledge of MOND comes, at present, from the study of purely gravitational systems (galactic systems, the solar system, etc.) it is still an open question how exactly to extend the third MOND tenet to systems involving arbitrary interactions. One possibility is to require that for $\\az\\rar\\infty$, the limiting equations of motion can be brought to a form where $G$, $\\az$, and $m_i$ appear only as $G\\az^2=\\az/\\mz$ and $m_i/\\az$. This requirement will automatically cause $\\az$, $G$, and $m_i$ to appear as $m_i/\\mz$ in the deep MOND limit of purely gravitational systems, as in such cases $G$ and $m_i$ always appear in equations as $Gm_i$. Such a general requirement would also replace Newton's second law $\\vF=m\\va$ by $\\vF=m\\vQ/\\az$ in the deep MOND regime, where $\\vQ$ is some functional with dimensions of acceleration squared that does not depend on $\\az$. \\par Detailed reviews of phenomenological and theoretical aspects of the paradigm can be found, for instance, in Sanders and McGaugh (2002), and, more recently, in Scarpa (2006) and in Bekenstein (2006). It follows from the third tenet that an underlying MOND theory must be nonlinear in the sense that an acceleration of a test particle due to a combination of several fields is not simply the sum of the accelerations produced by the individual fields. Take, as an example, the purely gravitational case where we modify the equations for the gravitational field. Linearity would mean that in the nonrelativistic limit of the theory the acceleration of a test particle at position $\\vr$ in the field of $N$ masses $m_i$ at positions $\\vr_i$ is given by \\beq \\va=\\sum_{i=1}^N m_i q_i(G, \\az, \\vr, \\vr_1,...,\\vr_N). \\eeqno{lino} The third assumption says that in the deep MOND case $q_i\\propto \\mz^{-1}$, but this is dimensionally impossible. Clearly, the acceleration produced even by a single point cannot be linear in its mass, as we shall see explicitly below. \\par It also follows from the third tenet, and the assumption that $\\az$ is the only new dimensioned constant, that the deep MOND limit of any theory must satisfy the following scaling laws for purely gravitational systems: In general, on dimensional grounds, all physics must remain the same under a change of units of length $\\ell\\rar\\lambda\\ell$, of time $t\\rar\\lambda t$, and no change in mass unit $m\\rar m$. Under these we have $\\az\\rar \\lambda^{-1}\\az$ and $G\\rar \\lambda G$; so, the constant $\\mz$, which alone appears in the limiting theory, is invariant under the scaling\\footnote{Other dimensioned quantities, such as the gravitational potential, have to be scaled appropriately. Note that velocities are invariant.}. This tells us that we are, in fact, exempt from scaling the constants of the theory when we scale $(\\vr,t)\\rar \\lambda(\\vr,t)$. The theory is thus invariant under this scaling; namely, if a certain configuration is a solution of the equations, so is the scaled configuration. Specific theories may have even higher symmetries; for example, the above scaling property is only part of the conformal invariance of the deep MOND limit for the particular MOND theory of Bekenstein \\& Milgrom (1984), as found in Milgrom (1997). \\par As a corollary of the scaling invariance we have: if $\\vr(t)$ is a trajectory of a point body in a configuration of masses $m_i$ at positions $\\vr_i(t)$ (which can be taken as fixed, for example), then $\\hat\\vr(t)=\\lambda\\vr(t/\\lambda)$ is a trajectory for the configuration where $m_i$ are at $\\lambda\\vr_i(t/\\lambda)$, and the velocities on that trajectory are $\\hat\\vV(t)=\\vV(t/\\lambda)$. (An extended mass changes its size and density such that the total mass remains the same. A point mass remains a point mass of the same value.) \\par Since $\\mz$ has dimensions of $mt^4/\\ell^4$, another scaling under which it is invariant is $m_i\\rar\\lambda m_i$, $\\vr_i\\rar\\vr_i$, $t\\rar\\lambda^{-1/4} t$. (The most general scaling allowed is $m\\rar\\lambda m$, $\\vr_i\\rar\\kappa\\vr_i$, $t\\rar\\kappa\\lambda^{-1/4}t$.) This means that for a purely gravitational system deep in the MOND regime, scaling all the masses leaves all trajectory paths the same but the bodies traverse them with all velocities scaling as $m^{1/4}$; accelerations then scale as $m^{1/2}$. \\par It is enlightening to draw an analogy between the role of $\\az$ in MOND with the role of $\\hbar$ in quantum physics, or that of $c$ in relativity. These constants, each in its own realm, mark the boundary between the classical and modified regimes; so formally pushing these boundaries to the appropriate limits ($c\\rar\\infty$, $\\hbar\\rar0$, $\\az\\rar 0$) one restores the corresponding classical theory (for quantum theory, in some weak sense). In addition, these constants enter strongly the physics in the modified regime where they feature in various phenomenological relations. For example, $\\hbar$ appears in the black body spectrum, the photoelectric effect, atomic spectra, and in the quantum Hall effect. The speed of light appears in the Doppler effect, in the mass vs. velocity relation, and in the radius of the Schwarzschild horizon. Without the respective, underlying theory, these disparate phenomena would appear totally unrelated, and the appearance of the same numerical constant in all of them would constitute a great mystery. The MOND paradigm similarly predicts a number of laws related to galactic motion, some of which are qualitative, but many of which are quantitative and involve $\\az$. Since they appear to be obeyed by nature it should indeed be a great mystery why that should be so without the underlying MOND paradigm (i.e. with Newtonian dynamics plus DM). I now discuss some of these predictions in some detail. ", "conclusions": "" }, "0801/0801.2303_arXiv.txt": { "abstract": "We present here some of the first results we have obtained on the study of the optical spectra of {\\em Spitzer/MIPS} 24 $\\rm \\mu m$-selected galaxies in the COSMOS field. This is part of a series of studies we are conducting to analyse the optical spectral properties of mid-infrared (mid-IR) galaxies with different IR luminosities up to high redshifts. The results shown here correspond to the brightest $S_{24 \\, \\rm \\mu m}>2 \\, \\rm mJy$ IR galaxy population at $z<1$. ", "introduction": "Since the discovery of an extragalactic component for the IR background a decade ago~\\cite{ref:pug}, numerous studies have demonstrated the key importance of IR galaxies to reconstruct the history of galaxy star formation and stellar mass assembly. Recent surveys conducted with the {\\em Spitzer Space Telescope} have allowed for the resolution of more than 70\\% of the mid and far-IR background light~\\cite{ref:dole} and a quite detailed understanding of the evolution of IR galaxies with cosmic epoch, e.g.~\\cite{ref:lefl,ref:cap06a,ref:cap06b,ref:cap07}. The IR galaxy luminosity function has a strong positive luminosity evolution from redshifts $z=0$ to 1 (see~\\cite{ref:lefl,ref:cap07}), implying that more luminous IR galaxies dominate the IR output with increasing redshift. Beyond $z\\sim1$, basically all the IR light is produced by luminous and ultra-luminous IR galaxies (LIRGs and ULIRGs, respectively) with bolometric IR luminosities $L_{\\rm bol.}> 10^{11} \\, \\rm L_\\odot$. Active galactic nuclei (AGN) have some significant (but not yet well-determined) contribution to the IR background and govern the shape of the bright-end of the mid-IR LF at $z\\sim2$~\\cite{ref:cap07}. The study of the optical spectra of IR galaxies can reveal several aspects of the physical conditions in which star formation and AGN activity take place. What is the evolutionary stage of the hosts of IR activity? What is the relationship between dust emission and the observed reddening in optical bands? Does the IR phase accompany all the process of star formation or is it set at a particular time? The large sample of zCOSMOS spectra~\\cite{ref:lil07} being taken over the 2 deg$^2$ COSMOS field offers a perfect opportunity to address these questions. ", "conclusions": "" }, "0801/0801.2245_arXiv.txt": { "abstract": "{} { We have carried out an accurate investigation of the Ca isotopic composition and stratification in the atmospheres of 23 magnetic chemically peculiar (Ap) stars of different temperature and magnetic field strength. } { With the UVES spectrograph at the 8\\,m ESO VLT, we have obtained high-resolution spectra of Ap stars in the wavelength range 3000--10000\\,\\AA. Using a detailed spectrum synthesis calculations, we have reproduced a variety of Ca lines in the optical and ultraviolet spectral regions, inferring the overall vertical distribution of Ca abundance, then we have deduced the relative isotopic composition and its dependence on height using the profile of the the IR-triplet \\ion{Ca}{ii} line at $\\lambda$~8498\\,\\AA. } { In 22 out of 23 studied stars, we found that Ca is strongly stratified, being usually overabundant by 1.0--1.5\\,dex below $\\log\\tau_{5000}\\approx -1$, and strongly depleted above $\\log\\tau_{5000}=-1.5$. The IR-triplet \\ion{Ca}{ii} line at $\\lambda$~8498\\,\\AA\\ reveals a significant contribution of the heavy isotopes $^{46}$Ca and $^{48}$Ca, which represent less than 1\\,\\% of the terrestrial Ca isotopic mixture. We confirm our previous finding that the presence of heavy Ca isotopes is generally anticorrelated with the magnetic field strength. Moreover, we discover that in Ap stars with relatively small surface magnetic fields ($\\le$\\,4--5 kG), the light isotope $^{40}$Ca is concentrated close to the photosphere, while the heavy isotopes are dominant in the outer atmospheric layers. This vertical isotopic separation, observed for the first time for any metal in a stellar atmosphere, disappears in stars with magnetic field strength above 6--7\\,kG. } { We suggest that the overall Ca stratification and depth-dependent isotopic anomaly observed in Ap stars may be attributed to a combined action of the radiatively-driven diffusion and light-induced drift. } ", "introduction": "\\label{intro} After the pioneering work by Michaud (\\cite{Michaud70}), atomic diffusion in stellar envelopes and atmospheres has been recognized as the main process responsible for the atmospheric abundance anomalies in the peculiar stars of the upper main sequence. Due to their unique characteristics, such as an extremely slow rotation, strong, global magnetic fields, and the absence of convective mixing, magnetic chemically peculiar (Ap and Bp) stars exhibit the most clear manifestation of diffusion effects and thus represent privileged laboratories for investigation of chemical transport processes and magnetohydrodynamics. Detailed diffusion calculations performed for a set of chemical elements in the atmospheres of magnetic peculiar stars predict separation of chemical elements over the stellar surface and with height in stellar atmosphere (abundance stratification). These theoretical predictions can be directly tested through the comparison with empirical maps of chemical elements inferred from observations. For a small number of elements, including Ca, an effect of the vertically stratified element distribution on the spectral line profiles was demonstrated in early studies (Borsenberger at al. \\cite{BPM81}). However, the absence of high-resolution, high signal-to-noise spectroscopic observations did not allow a robust comparison between observations and theoretical diffusion modelling. This step was carried out by Babel (\\cite{Babel92}), who calculated the Ca abundance distribution in the atmosphere of magnetic Ap star 53~Cam and showed that the unusual shape of \\ion{Ca}{ii} K line -- a combination of the wide wings and extremely narrow core (Babel \\cite{Babel94}; Cowley et al. \\cite{CHK06}) -- is a result of a step-like Ca distribution with abundance decrease at $\\log\\tau_{5000}\\approx-1$. Following Babel, the step-function approximation for the abundance distributions was employed in many stratification studies based on the observed profiles of spectral lines (\\txtbf{Savanov et al. \\cite{SKV01}}; Bagnulo et al. \\cite{bagetal01}; Wade et al. \\cite{WLRK03}; Ryabchikova et al. \\cite{RPK02,RLK05,RRKB06}; \\txtbf{Glagolevskii et al. \\cite{GRC05}; Cowley et al. \\cite{CHC07}}). Ca was found to be stratified in nearly the same way as in 53 Cam (enhanced concentration of Ca below $\\log\\tau_{5000}\\approx-1$ and its depletion above this level) in all stars for which stratification analysis have been performed: $\\beta$~CrB (Wade et al. \\cite{WLRK03}), $\\gamma$~Equ (Ryabchikova et al. \\cite{RPK02}), HD~204411 (Ryabchikova et al. \\cite{RLK05}), HD~133792 (Kochukhov et al. \\cite{KTRM06}) and HD~144897 (Ryabchikova et al. \\cite{RRKB06}). In addition to remarkable diffusion signatures, recently another Ca anomaly was detected, first in the spectra of the Hg-Mn stars by Castelli \\& Hubrig (\\cite{CastH04}) and then in Ap stars by Cowley \\& Hubrig (\\cite{CH05}). These authors found a displacement of the lines of \\ion{Ca}{ii} IR triplet due to significant contribution of the heavy Ca isotopes. Ryabchikova, Kochukhov \\& Bagnulo (see review paper by Ryabchikova (\\cite{MONS05})) were the first to apply spectrum synthesis calculations to investigate effects of Ca isotopes on the calcium line profile shape in magnetic CP stars. In particular, we have demonstrated the general anticorrelation between the presence of heavy Ca and magnetic field strength: in Ap stars the contribution of heavy Ca isotopes decreases with the increase of the field modulus, and disappears when the field strength exceeds $\\sim$\\,3\\,kG. Cowley et al. (\\cite{CHC07}) have come to a similar conclusion, but with reservations. In the present paper we summarise our detailed analysis of the vertical stratification of Ca abundance in the atmospheres of magnetic Ap stars of different temperatures and magnetic field strengths. We combine this vertical Ca mapping with the analysis of the Ca isotopic anomaly and its dependence on height, based on the spectrum synthesis modelling of the IR triplet \\ion{Ca}{ii} line at $\\lambda$~8498\\,\\AA. Our study is the first to present a homogeneous and systematic determination of the vertical stratification of a given element in the atmospheres of a large number of stars, enabling us to study the signatures of atmospheric atomic diffusion in the presence of strong magnetic field as a function of stellar parameters and magnetic field intensity. Our paper is organized as follows. In Sect.~\\ref{obs} we describe spectroscopic observations and the data reduction. Determination of the stellar atmospheric parameters is detailed in Sect.~\\ref{parameters}. Sects.~\\ref{synthesis}--\\ref{isot} present our methodological approach to magnetic spectrum synthesis, determination of Ca stratification and the study of Ca isotopic composition. Our findings are summarised in Sect.~\\ref{results} and are discussed in Sect.~\\ref{discus}. ", "conclusions": "\\label{discus} In this paper we have investigated the vertical stratification and isotopic anomaly of Ca in a large sample of cool magnetic Ap stars. Our study is the first to address the interesting problem of the presence of heavy Ca isotopes in chemically peculiar stars with detailed polarized radiative transfer calculations, which take into account the effects of the magnetic field and chemical separation in stellar atmospheres. We derive stratification of Ca for 23 Ap stars using a sample of \\ion{Ca}{i} and \\ion{Ca}{ii} lines distributed over a broad spectral range. All but one program stars clearly show signatures of the Ca stratification, whereas comparison stars reveal no inhomogeneities in the vertical Ca distribution when analysed with the same techniques and atomic data. Although our stratification modelling was based on a simplified step-function approximation and adopted a simplified homogeneous model for the magnetic field geometry, the inferred parameters of the vertical Ca distributions impose important observational constraints for theoretical models of the radiative diffusion processes in stellar atmospheres. Analysis of the IR Ca triplet line \\ion{Ca}{ii} 8498\\,\\AA\\ provided information on the relative contribution of different Ca isotopes. Spectrum synthesis calculations show that significant isotopic shifts observed in the core of \\ion{Ca}{ii} 8498\\,\\AA\\ cannot be attributed to the overabundance of heavy Ca isotopes throughout the whole stellar atmosphere. Instead, \\txtbf{we show that} the Ca line profile shape \\txtbf{is consistent with} the vertical separation of different Ca isotopes, with heavy Ca located in a cloud above the most abundant isotope $^{40}$Ca. Even though the presence of heavy Ca is prominent in the line core, the normal Ca isotope dominates the line wings and is more abundant in the lower atmosphere where the total Ca abundance is also much larger. Thus, our \\txtbf{tentative} model calls for a less extreme heavy Ca enrichment than was suspected in previous investigations limited to the centroid measurements of the \\ion{Ca}{ii} 8498\\,\\AA\\ line core. Recently different stratification of He isotopes was found by Bohlender (\\cite{B05}) in the analysis of $^{3}$He and related stars. That paper used a similar approach to modelling chemical stratification, but has only addressed He vertical distribution in hot non-magnetic chemically peculiar stars. We have successfully used the model with Ca stratification and vertical isotopic separation to explain the appearance of the \\ion{Ca}{ii} 8498\\,\\AA\\ in all stars showing excess of heavy Ca isotopes. The prominent anticorrelation between the presence of heavy Ca and magnetic field strength, first reported by Ryabchikova (\\cite{MONS05}), is confirmed and strengthened in the present paper. We find that only stars with sufficiently weak field show traces of heavy Ca. According to our knowledge, this interesting relation is the only case when definite dependence of the chemical abundance characteristic on the magnetic field strength is found for Ap stars. Furthermore, we find that in stars with stronger fields the heavy Ca isotopes tend to accumulate higher in the atmosphere. According to the results of our study, pulsating (roAp) and non-pulsating Ap stars are not distinguished by the characteristics of their Ca isotopic anomaly. Both groups of stars are equally likely to show an excess of heavy Ca and follow the same trend of the heavy isotope contribution versus the magnetic field strength and the \\ion{Ca}{ii} 8498\\,\\AA\\ line intensity. If the overall distribution of Ca abundance in the atmospheres of Ap stars follows the predictions of the radiatively driven diffusion, our results on the isotopic separation favour the light-induced drift (LID) as the main process responsible for this separation. According to Atutov \\& Shalagin (\\cite{AS88}), LID arises when the radiation field is anisotropic inside the line profile. Such an anisotropy takes place for a line of the trace isotope, for instance $^{46}$Ca and $^{48}$Ca in the terrestrial calcium mixture, which is located in the wing of a strong line of the main isotope $^{40}$Ca. The main isotope induces the drift velocity for other isotopes. If the trace isotope's line is located in the red wing of the line due to main isotope, the drift velocity is directed towards the upper atmosphere and the trace isotopes are pushed upwards. This is in agreement with observed vertical distribution of Ca isotopes. The Zeeman splitting changes the line shape and decreases the flux anisotropy for the line of trace isotope. When magnetic field becomes strong enough, $\\sim$5--6~kG, the flux anisotropy disappears and the isotopic separation is ceasing. Therefore, the observed Ca isotopic anomaly in magnetic stars may be qualitatively explained by the combined action of the radiatively-driven diffusion and the light-induced drift. Theoretic study of a combination of the radiatively-driven diffusion and LID was presented by LeBlanc \\& Michaud (\\cite{LM93}) for He. These authors showed that LID accelerates separation of $^{3}$He from $^{4}$He in hotter CP stars. Detailed theoretical chemical diffusion calculations (LeBlanc \\& Monin \\cite{LM04}) should incorporate LID in order to test our hypothesis that this effect may be important for the chemical transport processes in cool Ap-star atmospheres." }, "0801/0801.4769_arXiv.txt": { "abstract": "{We present results of a study of large-scale neutral hydrogen (\\HI) gas in nearby radio galaxies. We find that the early-type host galaxies of different types of radio sources (compact, \\FRI\\ and \\FRII) appear to contain fundamentally different large-scale \\HI\\ properties: enormous regular rotating disks and rings are present around the host galaxies of a significant fraction of low power compact radio sources, while no large-scale \\HI\\ is detected in low power, edge-darkened \\FRI\\ radio galaxies. Preliminary results of a study of nearby powerful, edge-brightened \\FRII\\ radio galaxies show that these systems generally contain significant amounts of large-scale \\HI, often distributed in tail- or bridge-like structures, indicative of a recent galaxy merger or collision. Our results suggest that different types of radio galaxies may have a different formation history, which could be related to a difference in the triggering mechanism of the radio source. If confirmed by larger studies with the next generation radio telescopes, this would be in agreement with previous optical studies that suggest that powerful \\FRII\\ radio sources are likely triggered by galaxy mergers and collisions, while the lower power \\FRI\\ sources are fed in other ways (e.g. through the accretion of hot IGM). The giant \\HI\\ disks/rings associated with some compact sources could - at least in some cases - be the relics of much more advanced mergers.} \\FullConference{From planets to dark energy: the modern radio universe\\\\ October 1-5 2007\\\\ University of Manchester, Manchester, UK} \\begin{document} ", "introduction": "Radio sources are generally hosted by early-type galaxies. A significant fraction of these radio galaxies, in particular the more powerful ones, show optical peculiarities (tail, bridges, shells, etc.) that are indicative of galaxy mergers or interactions \\citep{hec86,bau92}. This has led to the suggestion that galaxy mergers or collisions could be the trigger for the radio-AGN activity in these systems. Here we present the large-scale neutral hydrogen (\\HI) properties of different types of nearby radio galaxies in order to investigate whether these systems have been involved in a recent or past gas-rich merger or interaction. Simulation show that in a major merger between gas-rich galaxies, large-scale gaseous tidal structures can be expelled from the merging system, which after several rotations ($\\gtrsim$ Gyr) can partially fall back onto the host galaxy and settle in a low-density, regular rotating disk or ring \\citep{bar02}. \\HI\\ observations -- in particular when combined with a stellar population analysis of the host galaxy \\citep{tad05} -- therefore provide an excellent tool to trace and date galaxy mergers on relatively long time-scales, which can be compared with the age of the radio-loud AGN in these galaxies in order to investigate a possible connection with the triggering of the radio source. ", "conclusions": "\\label{sec:discussion} Our results of large-scale \\HI\\ around nearby radio galaxies imply that there could be {\\sl a fundamental difference in large-scale \\HI\\ content between the various types of radio galaxies}, which is visualized in more detail in Fig.~\\ref{fig:HImassplot}. While the host galaxies of several low power compact radio sources contain a large-scale regular rotating \\HI\\ disk or ring (possibly as the result of an advanced merger), the extended \\FRI\\ radio galaxies in the same sample lack any large-scale \\HI\\ structures outside the optical body of the host galaxy up to a conservative detection limit of a few $\\times 10^{8} M_{\\odot}$. On the other hand, preliminary results for a small sample of the more powerful \\FRII\\ radio sources show that most of their host galaxies do contain significant amounts of \\HI, often distributed in tail- or bridge-like structures. If confirmed by future studies with greater statistical significance and lower detection limits, this could mean that there is a fundamental difference in the formation history of different types of radio galaxies and, related, the triggering of their radio source. This would be in agreement with optical studies that suggest that powerful \\FRII\\ sources are generally fueled by major mergers, while \\FRI\\ source are likely powered in another way, for example through the accretion of gas from the hot inter-galactic medium \\citep{hec86,bau92}. Future instruments, with eventually the Square Kilometre Array, will be essential to verify our results by observing \\HI\\ emission in large statistical samples of radio galaxies. \\begin{figure}[t] \\centering \\includegraphics[width=10cm]{Proc_SKA_HImass_v3_lowres.eps} \\caption{Large-scale \\HI\\ mass plotted against the total linear extent of the radio source. Black circles and arrows represent the sources from our B2 sample (compact + \\FRI), while red squares and arrows are preliminary results from our \\FRII\\ sample. In case of non-detection a conservative upper limit (3$\\sigma$ across 400 \\kms) was estimated.} \\label{fig:HImassplot} \\end{figure}" }, "0801/0801.3463_arXiv.txt": { "abstract": "The 21-cm anisotropies from the neutral hydrogen distribution prior to the era of reionization is a sensitive probe of primordial non-Gaussianity. Unlike the case with cosmic microwave background, 21-cm anisotropies provide multi-redshift information with frequency selection and is not damped at arcminute angular scales. We discuss the angular trispectrum of the 21-cm background anisotropies and discuss how the trispectrum signal generated by the primordial non-Gaussianity can be measured with the three-to-one correlator and the corresponding angular power spectrum. We also discuss the separation of primordial non-Gaussian information in the trispectrum with that generated by the subsequent non-linear gravitational evolution of the density field. While with the angular bispectrum of 21-cm anisotropies one can limit the second order corrections to the primordial fluctuations below $f_{\\rm NL}\\sim 1$, using the trispectrum information we suggest that the third order coupling term, $f_2$ or $g_{\\rm NL}$, can be constrained to be arounde 10 with future 21-cm observations over the redshift interval of 50 to 100. ", "introduction": "The cosmic 21-cm background involving spin-flip line emission or absorption of neutral hydrogen contains unique signatures on how the neutral gas evolved from last scattering at $z \\sim 1100$ to complete reionization at $z < 10$ \\cite{Furlanetto}. Subsequent to recombination, the temperature of neutral gas is coupled to that of the cosmic microwave background (CMB). At redshifts below $\\sim$ 200 the gas cools adiabatically, its temperature drops below that of the CMB, and neutral hydrogen resonantly absorbs CMB flux through the spin-flip transition \\cite{field,loeb,Bharadwaj}. The inhomogeneous neutral hydrogen density distribution generates anisotropies in the brightness temperature measured relative to the blackbody CMB \\cite{zaldarriaga}. The large cosmological and astrophysical information content in 21-cm background is well understood in the literature \\cite{loeb,Iliev,Santos,Bowman,McQuinn,Pen,sigurdson}. Parallel to the large effort in analytical and numerical calculations of the 21-cm properties, there are now several first generation 21-cm experiments underway focusing on the 21-cm signal during the era of reionization. At the low redshifts probed by these first generation interferometers, the 21-cm signal is modified by the astrophysics of first sources and the associated UV photon background. There, one naturally expects fluctuations to be dominated by inhomogeneities in source properties \\cite{Zahn,Santos2}. The associated non-Gaussianity in 21-cm anisotropies leads to a measurable three-point correlation function or a bispectrum \\cite{Coo05,Bharadwaj2,morales2}. Such a non-Gaussianity is also expected to dominate the signature in 21-cm anisotropies generated by the primordial non-Gaussian density field. However, the 21-cm background generated by neutral hydrogen at redshifts of 50 to 100 prior to the onset of reionization and the appearance of first stars is expected to provide a cleaner probe of the primordial density perturbations in the same manner CMB observations are used to study primordial fluctuations \\cite{loeb}. If the primordial fluctuations are non-Gaussian, then the 21-cm anisotropies at these high redshifts will naturally contain a signature associated with that non-Gaussianity \\cite{Cooray,pillepich} The primordial non-Gaussianity in the density field can be studied with the three-point and higher-order correlation functions of the 21-cm background. In particular, the second order corrections to the density perturbations generated by primordial non-Gaussianity lead to a bispectrum with a dependence on the second-order correction to the curvature perturbations, $f_{\\rm NL}$ \\cite{Komatsu,Bartolo3}. With future low frequency data out to $z \\sim 100$, 21-cm background anisotropies could potentially limit $f_{\\rm NL} < 0.1$ \\cite{Cooray}. In comparison, the expected non-Gaussianity under standard inflation is of order $|n_s-1|$ and with the scalar spectral index $n_s \\sim 0.98$ \\cite{Spergel}, $f_{\\rm NL}$ is expected to be well below unity \\cite{Maldacena}. The primordial non-Gaussianity parameter at the second order $f_{\\rm NL}$, however, has a correction associated with evolution of second and higher-order perturbations after inflation \\cite{Bartolo}. For standard slow-roll inflation then $f_{\\rm NL} = -5/12(n_s-1) + 5/6 +3/10f(k)$ where the last term is momentum dependent. In this case $f_{\\rm NL}$ is at the level of a few tenths and could be as high as 1. The ability for 21-cm anisotropies to probe $f_{\\rm NL}$ as low as 0.1 is important since even a perfect CMB experiment limited by cosmic variance alone can only restrict $f_{\\rm NL} > 3$ \\cite{Komatsu,Babich,Smith} while there is no significant improvement when using low redshift large-scale structure \\cite{Scoc,Dalal}. The possibility to make primordial non-Gaussianity measurements with the 21-cm background is important since compared to most other probes of inflationary parameters, $f_{\\rm NL}$ is one of the few parameters for which we have limited number of probes sensitive to the low amplitude non-Gaussianity expected under standard slow-roll inflation. While CMB as a probe of non-Gaussianity is well known, when compared to CMB temperature and polarization anisotropies, 21-cm background has two distinct advantages: (1) the ability to probe multiple redshifts based on frequency selection and (2) the lack of a damping tail in the 21-cm anisotropy spectrum, unlike damping of CMB anisotropies at a multipole around 2000. While the 21-cm background as a potentially interesting and a useful probe of $f_{\\rm NL}$ is now known, different scenarios for primordial fluctuations may produce a small $f_{\\rm NL}$ but a large third-order correction. In certain alternative models of inflation higher order terms may be significant even if the second-order term is small and such scenarios include the new ekpyrotic cosmology \\cite{Buchbinder} and, under certain conditions, the curvaton model \\cite{Sasaki}. Thus, beyond the non-Gaussianity at the three-point level with the bispectrum, it is also useful to study the non-Gaussianities of the 21-cm background at the four point-level involving the trispectrum. Here we show that the angular trispectrum of 21-cm anisotropies contains a measurable non-Gaussianity from primordial fluctuations if the scale-independent cubic corrections to the gravitational potential captured with an amplitude parameter $f_2$ (or $g_{\\rm NL}$ as described in Ref.~\\cite{Amico}) has a value of order $\\sim 10$, even if the scale-independent quadratic corrections to the gravitational potential captured by $f_{\\rm NL}$ has a value around $\\sim$ 1. While $f_{\\rm NL}$ is currently constrained with WMAP data \\cite{Spergel}, there is no real constraint on this third order non-Gaussianity parameter. There are, however, theoretical expectations: for slow-roll inflation, a trispectrum of the form $T_4=1/2\\tau_{\\rm NL} [P(k_1)P(k_2)P(k_3) + ...]$ is expected for curvature perturbations \\cite{Seery} with an expectation value of $\\tau \\lesssim r/50$ where $r$ is the tensor-to-scalar ratio ($r \\lesssim 0.6$ is recent CMB data \\cite{Spergel}). In such a model, there is also a direct connection between $\\tau_{\\rm NL}$ and $f_{\\rm NL}$ such that $\\tau_{\\rm NL}=(6f_{\\rm NL}/5)^2$ (see the discussion in Ref.~\\cite{Kogo:2006} for connections between coupling terms of various models for primordial trispectrum). While such a relation is generally assumed when constraining $\\tau_{\\rm NL}$ \\cite{Lyth}, in future it may be that one can directly test the above relation between $f_{\\rm NL}$ and $\\tau_{\\rm NL}$ with data. To measure the non-Gaussianity at the four-point level, we introduce the three-to-one correlator statistic, extending the two-to-one correlator of Ref.~\\cite{Coo01} and applied to WMAP data in Ref.~\\cite{Szapudi}. We optimize the angular power spectrum of the 3-1 correlator to detect the primordial trispectrum by appropriately filtering 21-cm anisotropy data to remove the non-Gaussian confusion generated by the non-linear evolution of gravitational perturbations. To do this properly, one requires prior knowledge on the configuration dependence of the primordial 21-cm trispectrum, but its amplitude is a free variable to be determined from the data. This paper is organized as following: we first discuss the bispectrum in 21-cm anisotropies associated with primordial perturbations resulting from quadratic corrections to the primordial potential. We discuss ways to measure this bispectrum in the presence of other non-Gaussian signals and determine the extent to which $f_{\\rm NL}$ can be measured from 21-cm background data. In the numerical calculations described later, we take a fiducial flat-$\\Lambda$CDM cosmological model with $\\Omega_b=0.0418$, $\\Omega_m=0.24$, $h=0.73$, $\\tau=0.092$, $n_s=0.958$, and $A(k_0 = 0.05\\ \\mathrm{Mpc}^{-1})=2.3\\times10^{-9}$. This model is consistent with recent measurements from WMAP \\cite{Spergel}. ", "conclusions": "\\label{sec:results} In Fig.~1 we summarize the power-spectrum of 21-cm anisotropies generated by the neutral hydrogen distribution at a redshift of 100 with a bandwidth for observations of 1 MHz. Here, we also plot $X_l^{\\rm prim}$, $X_l^{\\rm grav}$, and $N_l^{\\rm tot}$ for the same redshift with the optimal filter applied with $f_{\\rm NL}=10$ and $f_2=1$ as the non-Gaussian scale-independent amplitude of the primordial second- and third-order curvature perturbations, respectively. As shown, $N_l^{\\rm tot} > X_l^{\\rm grav}$, suggesting that the noise term is dominated by the Gaussian variance (second term in equation~\\ref{eqn:nl}). This statement is independent of $f_{\\rm NL}$ and $f_2$ and thus the non-Gaussian detection is dominated by the Gaussian term in $N_l$ regardless of what is assumed about non-Gaussianity. Note that we have estimated $N_l^{\\rm tot}$ in Fig.~1 under the assumption that observations are limited only by the cosmic variance and not accounting for any instrumental noise variance, which will also lead to a cut-off in $l$ out to which we can make measurements. Using the cosmic variance alone allows us to establish the potentially achievable limit and compare directly with cosmic variance limit with CMB data. When calculating $X_l^{\\rm prim}$ and $X_l^{\\rm grav}$ in equation~(\\ref{eqn:xl}) we set maximum value of $L$ in the sum associated with $T^{{l_1l_2}_{l_3l}}(L)$ to be $L_{\\rm max}=100$ We tested our calculation for $L_{\\rm max}=150$ and found results to be within a percent, but such a higher value slows the numerical calculation significantly. Finally, due to computational limitations of the numerical calculation, we restrict estimate of $X_l$ to $l=10^3$. In Fig.~1 when calculating $X_l^{\\rm grav}$, following the derivation in the Appendix and the discussion there, we use the exact analytical result for the 2-2 trispectrum of the density field with the mode coupling captured by $F_2(\\bk_1,\\bk_2)$ \\cite{Goroff,fry84}. For the 3-1 trispectrum of the density field under non-linear gravitational evolution, given that an analytical result for the angular trispectrum is cumbersome, we use the angular averaged value for $F_3(\\bk_1,\\bk_2,\\bk_3)$. We refer the reader to the Appendix for details. In Fig.~2 we summarize the estimate related to signal-to-noise ratio for a detection of $X_l^{\\rm prim}$ as a function of $l$. The typical signal-to-noise ratio, when measurements are out to a multipole of $10^3$, is at the level of $\\sim$ 0.5 if one assumes that the coupling parameters $f_{\\rm NL}=10$ and $f_2=3$ for 21-cm observations centered at a redshift of 100 over a bandwidth of 1 MHz. The signal-to-noise ratio for the case with $f_{\\rm NL}=0$ and $f_2=3$ is $\\sim$ 0.03. Since $S/N \\propto f_2$ when $f_{\\rm NL}=0$, out to $l=10^3$, a signal-to-noise ratio of 1 is achieved if $f_2 \\sim 10^2$. While there are neither strong theoretical motivations on the expected value of $f_2$ nor a real bound on its value from existing data, it is likely that with 21-cm data one can constrain $f_2$ down to a level well below this value for a single redshift. This is due to the fact that 21-cm observations lead to measurements at multiple redshifts, though one cannot make arbitrarily small bandwidths to improve the detection since at scales below a few Mpc, anisotropies in one redshift bin will be correlated with those in adjacent bins \\cite{Santos3}. For example, if 21-cm observations are separated to 30 independent bins over the redshift interval 50 to 100 (as can be achieved with 1 MHz bandwidths), then an approximate estimate of the cumulative signal-to-noise ratio, ${\\rm S}/{\\rm N} = \\sqrt{\\sum_z [{\\rm S}/{\\rm N}(z)]^2}$, is $\\sim 0.15 (f_{2}/3)$ if $f_{\\rm NL}=0$. In return, one can potentially probe $f_{\\rm NL}$ values as low as 20 roughly out to $l_{\\rm max}\\sim 10^3$. With the first-generation radio interferometers, we would at most survey 1\\% of the sky. Assuming instrument noise is dominating at multipoles above $10^3$ between 30 MHz and 60 MHz at 1 MHz bandwidths (corresponding to redshifts 30 to 100), we find a signal-to-noise ratio of $5 \\times 10^{-3} f_{2}$, which could lead to a limit on $f_2$ of order 200. Above discussion on the application of $X_l^{\\rm prim}$ assumes that $f_{\\rm NL}=0$. Since $X_l^{\\rm prim} \\propto f_{\\rm NL}^2$, if $f_{\\rm NL}$ is greater than one, the overall signal-to-noise ratio for the detection of the three-to-one angular power spectrum is increased and the dominant contribution to $X_l^{\\rm prim}$ comes from the coupling associated with $f_{\\rm NL}$ and not $f_2$. In the case when $f_2=0$, the $f_{\\rm NL}$ one probes with the trispectrum can be related to $\\tau_{\\rm NL}$ of Refs.~\\cite{Seery,Lyth} for the primordial trispectrum. In Fig.~2, we show the signal-to-noise ratio with $f_{\\rm NL}=10$ and $f_2=3$. Since in this case $f_{\\rm NL}$ term dominates, this provides an approximate estimate of the signal-to-noise ratio for $f_{\\rm NL}$ with the trispectrum. Using a single redshift bin out to $l_{\\rm max} \\sim 10^3$, 21-cm observations achieve a signal-to-noise ratio of 1 if $f_{\\rm NL} \\sim 15$. This in return constraints $tau_{\\rm NL} \\lesssim 300$. Assuming 30 redshift bins over the redshift interval 50 to 100, assuming $f_2=0$, we find that one can constrain $\\tau_{\\rm NL} < 50$. This result is only out to $l_{\\rm max} =10^3$, but since 21-cm observations are not damped as in the case of CMB observations, higher resolution data can improve limits on both $f_2$ and $\\tau_{\\rm NL}$ significantly especially if observations can be pushed to $l > 10^4$. While a detection of the CMB angular trispectrum has been motivated as a way to constrain $f_{\\rm NL}$ or $\\tau_{\\rm NL}$ \\cite{Kogo:2006}, this is probably not necessary with 21-cm data. Once radio interferometers start probing the redshift interval of 50 to 100, the angular bispectrum, which can be probed with the two-to-one angular power spectrum \\cite{Cooray}, can limit $f_{\\rm NL} < 0.1$. This will facilitate a separation of the contribution to the three-to-one power spectrum from $f_{\\rm NL}$ and $f_2$. While we have assumed that $f_{\\rm NL}$ and $f_2$ are momentum independent, it is likely that for specific models of inflation, these coupling terms are momentum dependent \\cite{Bartolo2} and then the ability to separate $f_{\\rm NL}$ and $f_2$ by combining 21-cm bispectrum and trispectrum information with the two-to-one and three-to-one correlator respectively will strongly depend on the exact momentum dependence of the coupling factors. We leave such a study for future research. While measuring $f_2$ is well motivated, one can also test the consistency between $f_{\\rm NL}$ probed by the bispectrum and the $\\tau_{\\rm NL}$ from the trispectrum related to the slow-roll inflation predictions for the non-Gaussianity. While we have not discussed the extent to which this consistency relation can be established with upcoming experiments after taking into account of instrumental noise it will also be useful to return to such a calculation in the future once 21-cm observations begin to probe the universe at $z > 10$. \\begin{figure}[t] \\includegraphics[scale=0.5,angle=-90]{Fig2.eps} \\caption{ Signal-to-noise ratio for a detection of $X_l^{\\rm prim}$ with $f_{\\rm NL}=10$ and $f_2=3$ (solid line) and $f_{\\rm NL}=0$ and $f_2=3$ (dot-dashed line). For reference, we show the signal-to-noise ratio associated with the detection of the full 21-cm bispectrum with a dashed line. } \\label{sn} \\end{figure} The 21-cm anisotropies from the neutral hydrogen distribution prior to the era of reionization is expected to be more sensitive to primordial non-Gaussianity than the cosmic microwave background due to both the three-dimensional nature of the 21-cm signal and the lack of a damping tail at arcminute angular scales. Previous calculations have discussed the extent to which 21-cm bispectrum can be used as a probe of primordial non-Gaussianity at the two-point level with an non-Gaussianity parameter $f_{\\rm NL}$ \\cite{Cooray,pillepich}. Here, we extend these calculations to discuss the possibility to use a four point statistic of the 21-cm background as a probe of the primordial non-Gaussianity associated with the trispectrum. We have calculated the angular trispectrum of the 21-cm background anisotropies and have introduced th three-to-one correlator and the corresponding angular power spectrum as a probe of the primordial trispectrum captured by both $f_{\\rm NL}$ and the third order non-Gaussianity parameter $f_2$ (described in some publications as $g_{\\rm NL}$). Since the primordial non-Gaussianity is confused with a non-Gaussian signal in the 21-cm background generated by the non-linear evolution of the density perturbations under gravitational evolution, we have discussed way to separate the two using an optimal filter. While with the angular bispectrum of 21-cm anisotropies one can limit the second order corrections to the primordial fluctuations as low as $f_{\\rm NL}\\sim 0.1$ below the value of $\\sim 1$ expected for inflationary models, using the trispectrum information we suggest that one can constrain third order coupling term $f_2$ to about few tens. If $f_{\\rm NL}$ is large, it could potentially be possible to test the consistency between $f_{\\rm NL}$ from the bispectrum and the slow-roll non-Gaussianity $\\tau_{\\rm NL}$ at the four-point level with the relation $\\tau_{\\rm NL}=(6f_{\\rm NL}/5)^2$. We hope to return to such a detailed study in the future." }, "0801/0801.4106.txt": { "abstract": "} }% %BeginExpansion \\begin{abstract} %EndExpansion The ratio of the Bondi and Jeans lengths is used to develop a cloud-accretion model that describes both an inner Bondi-type regime where gas pressure is balanced by the gravity of a central star and an outer Jeans-type regime where gas pressure is balanced by gas self-gravity. The gas density profile provided by this model makes a smooth transition from a wind-type inner solution to a Bonnor-Ebert type outer solution. It is shown that high-velocity dust impinging on this cloud will tend to pile-up due to having a different velocity profile than gas so that the dust-to-gas ratio is substantially enriched above the 1\\% ISM\\ level. %TCIMACRO{\\TeXButton{End abstract}{ ", "introduction": "Laboratory-scale plasma jets have been produced by a magnetohydrodynamic mechanism believed analogous to the mechanism responsible for driving astrophysical jets %TCIMACRO{\\TeXButton{\\citet{Hsu2002,Bellan2005}}{\\citep{Hsu2002,Bellan2005}}}% %BeginExpansion \\citep{Hsu2002,Bellan2005}% %EndExpansion . The laboratory jets are driven by capacitor bank power supplies that provide poloidal and toroidal magnetic fields and the jet acceleration mechanism can be considered as due to the pressure of the toroidal magnetic field inflating flux surfaces associated with the poloidal magnetic field. If these jets are indeed related to astrophysical jets, the obvious question arises as to what constitutes the power supply responsible for the toroidal and poloidal magnetic fields in the astrophysical situation. Existing models of astrophysical jets typically assume the poloidal field is simply given and that the toroidal field results from the rotation of an accretion disk twisting up the assumed poloidal field. The author has developed an alternate model which postulates that the toroidal and poloidal field result instead from a dusty plasma dynamo mechanism that converts the gravitational energy of infalling dust grains into an electrical power source that drives poloidal and toroidal electric currents creating the respective toroidal and poloidal fields. A brief outline of how infalling charged dust can drive poloidal currents has been presented in %TCIMACRO{\\TeXButton{\\citet{Bellan2007}}{\\citet{Bellan2007}}}% %BeginExpansion \\citet{Bellan2007}% %EndExpansion . An important requirement of the model is that there should be sufficient infalling dust to provide the jet power. It has been well established that the dust-to-gas mass ratio in the Interstellar Medium (ISM) is 1\\%. If one assumes, as has been traditional, that this ratio holds throughout the accretion process, then the gravitational energy available from infalling dust would be insufficient. However, %TCIMACRO{\\TeXButton{\\citet{Fukue2001}}{\\citet{Fukue2001}} }% %BeginExpansion \\citet{Fukue2001} %EndExpansion has recently shown via numerical solution of coupled dust and gas equations of motion that the dust-to-gas ratio can become substantially enriched during Bondi-type accretion. Star formation has also been previously examined from a molecular cloud physics point of view which differs significantly from the Bondi accretion point of view in the treatment of self-gravity and inflows. Molecular cloud physics is clearly important because observations indicate that stars form within the dense cores of molecular clouds (e.g., see %TCIMACRO{\\TeXButton{\\citet{Evans2001}}{\\citet{Evans2001}}}% %BeginExpansion \\citet{Evans2001}% %EndExpansion ). Analysis of the force balance in molecular clouds shows that the cloud radial density profile can be characterized by the Bonnor-Ebert sphere solution %TCIMACRO{\\TeXButton{\\citet{Ebert1955,Bonnor1956}}{\\citep{Ebert1955,Bonnor1956}% %}}% %BeginExpansion \\citep{Ebert1955,Bonnor1956}% %EndExpansion . This solution does not take into account flows associated with accretion. Dust dynamics is also not taken into account; instead it is typically assumed that the dust-to-gas mass ratio is fixed at 1\\% for all radii. To summarize the above discussion, we note that the Bondi-type analysis reported by %TCIMACRO{\\TeXButton{\\citet{Fukue2001}}{\\citet{Fukue2001}} }% %BeginExpansion \\citet{Fukue2001} %EndExpansion emphasizes inflow physics and dust-gas coupling but does not take into account self-gravity whereas molecular cloud analysis such as used by %TCIMACRO{\\TeXButton{\\citet{Evans2001}}{\\citet{Evans2001}} }% %BeginExpansion \\citet{Evans2001} %EndExpansion emphasizes self-gravity, but does not take into account inflow or dust-gas coupling. This paper will address the physics of coupled dust and gas accretion using a methodology similar in concept to that presented by %TCIMACRO{\\TeXButton{\\citet{Fukue2001}}{\\citet{Fukue2001}}}% %BeginExpansion \\citet{Fukue2001}% %EndExpansion , but extended to bridge the gap between Bondi accretion models and molecular cloud force balance models. \\ The enrichment mechanism observed by %TCIMACRO{\\TeXButton{\\citet{Fukue2001}}{\\citet{Fukue2001}} }% %BeginExpansion \\citet{Fukue2001} %EndExpansion will be examined in detail and will be shown to result from the inherently non-uniform nature of the Bondi/molecular cloud/ISM system. Properties of dust and gas for radii ranging from the cloud-ISM interface to the inner Bondi region will be considered by examining a sequence of successively smaller concentric regions. The basic character of each region and its scale, defined in terms of the nominal radial distance from a central object star, is as follows: \\begin{description} \\item \\textit{ISM scale:} The outermost scale is that of the Interstellar Medium (ISM). The ISM\\ has a gas density $\\sim$10$^{7}$ m$^{-3}$, a dust-to-gas mass ratio of 1 percent, a gas temperature $T_{g}^{ISM}\\sim100$ K$,$ and is optically thin. The ISM is assumed to be spatially uniform and to bound a molecular cloud having radius $r_{edge}$. \\item \\textit{Molecular cloud scale:}\\ The molecular cloud scale has much higher density than the ISM and is characterized by force balance between gas self-gravity\\ and gas pressure. The molecular cloud scale is sub-divided into a large, radially non-uniform low-density outer region and a small, approximately uniform, high-density inner core region. Clouds have a characteristic scale given by the Jeans length $r_{J}.$ The radial dependence of density is provided by the Bonnor-Ebert sphere solution which acts as the outer boundary of the Bondi accretion scale. \\item \\textit{Bondi accretion scale:} The Bondi accretion scale %TCIMACRO{\\TeXButton{\\citep{Bondi1952}}{\\citep{Bondi1952}} }% %BeginExpansion \\citep{Bondi1952} %EndExpansion is $\\sim r_{B}$ which is sufficiently small that gas self-gravity no longer matters so equilibrium is instead obtained by force balance between gas pressure and the gravity of a central object assumed to be a star having mass $M_{\\ast}$. The Bondi scale is sub-divided into three concentric radial regions: an outermost region where the gas flow is subsonic, a critical transition radius at exactly $r_{B}$ where the flow is sonic, and an innermost region where the gas flow is free-falling and supersonic. \\item \\textit{Collisionless dusty plasma scale (to be considered in a future publication): } Free-falling dust grains collide with each other in one of the above scales and coagulate to form large-radius grains which are collisionless and optically thin. The optically thin dust absorbs UV photons from the star, photo-emits electrons and becomes electrically charged. The charged dust grains are subject to electromagnetic forces in addition to gravity. Motions of charged dust grains relative to electrons result in electric currents with associated poloidal and toroidal magnetic fields [see preliminary discussion in %TCIMACRO{\\TeXButton{\\citet{Bellan2007}}{\\citet{Bellan2007}}}% %BeginExpansion \\citet{Bellan2007}% %EndExpansion ]. \\ \\ \\item \\textit{Jet scale (to be considered in a future publication): }The electric currents interact with the magnetic fields to produce magnetohydrodynamic forces which drive astrophysical jets in a manner analogous to that reported in \\ %TCIMACRO{\\TeXButton{\\citet{Hsu2002}}{\\citet{Hsu2002}} }% %BeginExpansion \\citet{Hsu2002} %EndExpansion and %TCIMACRO{\\TeXButton{\\citet{Bellan2005}}{\\citet{Bellan2005}}}% %BeginExpansion \\citet{Bellan2005}% %EndExpansion . \\end{description} \\bigskip The separation-of-scales requirement $r_{J}>>r_{B}$ implies existence of a small parameter% \\begin{equation} \\varepsilon_{BJ}\\ =\\frac{r_{B}}{r_{J}}\\ \\label{eps}% \\end{equation} quantifying the separation between the Bondi and Jeans scales. For purposes of relating the Bondi and Jeans scales to each other it is convenient to introduce a geometric-mean scale with characteristic length given by \\begin{equation} r_{gm}=\\sqrt{r_{B}r_{J}}=\\sqrt{\\varepsilon_{BJ}}r_{J}\\ \\label{def rgm}% \\end{equation} in which case% \\begin{equation} r_{B}\\ll r_{gm}\\ll r_{J}. \\label{separation}% \\end{equation} Gas and dust must be considered separately in each scale. We assume that gas motion influences dust motion but not vice-versa so that the dust dynamics can be ascertained after gas dynamics has been determined. This assumption is appropriate so long as dust mass and energy densities are small compared to corresponding gas densities or if the dust is decoupled from the gas. The sequence \\ of scales is summarized in Tables \\ref{gas table} and \\ref{dust table}. The numerical values of table elements with asterisks will be predicted by the model to be presented here while table elements with filled-in numbers represent\\ prescribed physical boundary conditions. Here $m_{g}$ is the gas molecular mass, $n_{g}$ is the gas density, $T_{g}$ is the gas temperature, $u_{g,d}$ are the radially inward gas and dust fluid velocities, and $\\ $ $\\rho_{g,d}$ are the gas and dust mass densities. Quantities with superscripts `ISM' are evaluated in the ISM, un-superscripted quantities refer to Bondi or molecular cloud regions. Our methodology will be conceptually similar to %TCIMACRO{\\TeXButton{\\citet{Fukue2001}}{\\citet{Fukue2001}} }% %BeginExpansion \\citet{Fukue2001} %EndExpansion but will also differ in important ways. %TCIMACRO{\\TeXButton{\\citet{Fukue2001}}{\\citet{Fukue2001}} }% %BeginExpansion \\citet{Fukue2001} %EndExpansion used a wind equation scheme to model Bondi-type physics and, unlike the analysis to be presented here, assumed the Bondi accretion region was directly bounded by the ISM, i.e., no molecular cloud region with Jeans-type scaling was taken into account. In addition, Fukue assumed that (i) the gas was heated by the combined effects of adiabatic compression and friction due to dust-gas collisions, and (ii) the entire region was optically thin so that the dust was subject to radiation pressure. Because of the complexity introduced by the heating of the gas, Fukue obtained results via numerical solution of three coupled differential equations (gas momentum, dust momentum, gas heating). Our approach will differ by assuming that (i) the gas is isothermal, (ii) a molecular cloud region lies between the Bondi accretion region and the ISM, and (iii) the system is optically thick outside the inner part of the Bondi region so that dust is shielded from optical radiation until it penetrates to the inner part of the Bondi region. Furthermore, rather than using numerical solutions, we will attempt analytic solutions as much as possible, a goal made feasible by the isothermal assumption. We believe our assumptions (i)-(iii) reasonably correspond to observations that gas is nearly isothermal and that stars form inside the cores of optically thick molecular clouds. \\bigskip% %TCIMACRO{\\TeXButton{B}{\\begin{table}[tbp] \\centering}}% %BeginExpansion \\begin{table}[tbp] \\centering %EndExpansion% \\begin{tabular} [c]{lllll}\\hline\\hline region & location & $n_{g}$ & $T_{g}$ & $u_{g}$\\\\\\hline\\hline ISM & $r>r_{edge}$ & 10$^{7}$ m$^{-3}$ & 100 K & *\\\\ B-E sphere & $r_{gm}\\sqrt{\\kappa T_{g}/m_{g}}$% \\end{tabular} \\caption{Sequence of regions for gas, * indicates quantity to be discussed/calculated in text}\\label{gas table}% %TCIMACRO{\\TeXButton{E}{\\end{table}}}% %BeginExpansion \\end{table}% %EndExpansion % %TCIMACRO{\\TeXButton{B}{\\begin{table}[tbp] \\centering}}% %BeginExpansion \\begin{table}[tbp] \\centering %EndExpansion% \\begin{tabular} [c]{llll}\\hline\\hline region & location & $\\rho_{d}/\\rho_{g}$ & $u_{d}$\\\\\\hline\\hline ISM & $r>r_{edge}$ & 0.01 & 3 km/s (turbulent acceleration)\\\\ B-E sphere & $\\sqrt{r_{B}r_{J}}10^{12}\\Lsun$), and suggests that our simple selection criterion effectively identifies a significant fraction of $z\\sim 2$ ULIRGs. This IRLD is also $\\approx26\\pm14$\\% of the total contributed by all $z\\sim2$ galaxies, and comparable to that contributed by the luminous UV-bright star-forming galaxy populations at $z\\approx 2$. We suggest that these DOGs are the progenitors of luminous ($\\sim 4L^*$) present-day galaxies and are undergoing an extremely luminous, short-lived phase of both bulge and black hole growth. They may represent a brief evolutionary phase between sub-millimeter-selected galaxies and less obscured quasars or galaxies. ", "introduction": "Infrared-luminous sources (with $L_{\\rm IR}>10^{11}{\\rm L_\\odot}$) dominate the bright end of the bolometric luminosity function of galaxies in the local universe \\citep[]{soi1986}. Their role in the story of galaxy evolution is not yet understood, but it is likely that their prodigious luminosities are the product of an extremely active phase during which these systems form stars and/or grow their central black holes at a rapid rate \\citep[]{san1988}. Did all large galaxies exhibit such a phase in their early formative stages, or is this active phase characteristic of only some of the most massive systems? To address these questions, we need to find and study the primary high-redshift galaxy populations contributing to the mid-IR and far-IR counts and understand their relation to the local galaxy population. We know a great deal about the local population of IR-luminous galaxies, due in large part to the pioneering observations made with the {\\it Infrared Astronomical Satellite} \\citep[{\\it IRAS}; e.g.,][]{soi1987a,san1988,san1996}. Their mid-IR number counts show strong evolution \\citep[e.g.,][]{elb1999,chary2001,pap2004,chary2004,marl2004}. Although rare locally, infrared luminous galaxies become an increasing fraction of the galaxy population at high redshift, dominating the IR energy density at redshifts $z\\sim1$ \\citep[e.g.,][]{fra2001,lag2004,gru2005,lef2005,per2005}. Far less is known about their higher redshift counterparts despite several ground-breaking studies. Prior to the launch of the {\\it Spitzer Space Telescope}, our knowledge of high-$z$ IR-luminous galaxies was derived from studies of a handful of the most extreme {\\it IRAS} sources, constraints on the redshift evolution out to $z\\sim 1$ \\citep[largely due to {\\it ISO}; e.g.,][]{fra2001,chary2001,lag2003}, small samples of galaxies discovered at sub-millimeter wavelengths \\citep[e.g.,][]{sma2002,bla2002,cha2004b,cha2005}, and several galaxies discovered as a result of their extreme optical-to-near-infrared colors \\citep[the dusty `Extremely Red Objects', or EROs, with $R-K>6$; e.g.,][]{hu1994,dey1999,mcc2004a}. The launch of the {\\it Spitzer Space Telescope} has revolutionized studies of the dusty galaxy population. It has provided the first mid-IR spectroscopy of dust emission in high-redshift star-forming galaxies and AGN, and revealed populations of galaxies that are extremely faint at (observed) visible wavelengths \\citep[e.g.,][Higdon et al. and Desai et al., both in preparation]{hou2005,yan2005,yan2007}. Many of these galaxies lie at redshifts $z\\simgt 1$, and their mid-IR spectra reveal either strong silicate absorption features, suggestive of an AGN, or emission from polycyclic aromatic hydrocarbons (PAHs) which generally arise in photodissociation regions associated with star-forming molecular clouds. A significant fraction of these active high-redshift galaxies are under-luminous at rest-frame UV wavelengths, suggesting that the prodigious star-formation and / or AGN accretion luminosity is hidden by dust that reprocesses and radiates the bulk of the energy at far-IR wavelengths \\citep[e.g.,][]{san1996}. In this paper, we present a simple and economical method for selecting these high-redshift dusty populations, based solely on the ratio of the observed 24$\\mu$m to optical flux ratio. This selection, applied to a wide area survey, results in a large, significant population of extremely luminous high-redshift galaxies for which there are no local counterparts. These are systems with unusually red spectral energy distributions, which we interpret as being the result of extinction by large columns of dust. These galaxies appear to be largely absent from (rest-frame) UV-selected high-redshift galaxy samples. We have measured a large number of spectroscopic redshifts for this population, and confirm that they lie at redshifts $z\\sim2$. Given their large mid-infrared luminosities and implied far-infrared luminosities, these dust obscured galaxies may be undergoing an intense burst of star formation, accretion activity (onto a nuclear black hole) or plausibly both. We speculate that this population of objects, which may be selected simply on the basis of optical to mid-infrared flux ratio, are good candidates for systems in the throes of formation of their stellar spheroids and nuclear black holes. Despite their rarity, this population contributes a significant fraction of the infrared luminosity density at $z\\sim 2$. This paper is structured as follows. In \\S2, we briefly describe the observations from which the sample is drawn and discuss the selection criteria we use to isolate the candidate high-redshift dust-obscured galaxy (hereinafter DOG) population. In \\S3, we present the main results: the optical-to-mid-infrared color distribution of the DOGs, their spectral energy distributions, redshift distribution and rest-frame UV and IR luminosities. In \\S4, we discuss the nature of the DOG population, derive their space density and their contribution to the $z\\approx 2$ infrared luminosity density, and discuss their possible relationship to other $z\\approx 2$ galaxy populations. We summarize our findings in \\S5. Throughout this paper we adopt a cosmology with ${\\rm H_0 = 70~km\\ s^{-1}\\ Mpc^{-1}}$, $\\Omega_m = 0.3$, $\\Omega_\\lambda = 0.7$. Magnitudes quoted in this paper are in Vega units, except where explicitly defined otherwise (e.g., in the discussion of rest-frame 2200\\AA\\ luminosities). ", "conclusions": "\\subsection{The Nature of the Extreme Red Population} What are the extreme red objects that become an increasing fraction of the population at fainter 24$\\mu$m flux densities? The spectroscopically measured redshifts for this population suggest prodigious luminosites ($10^{12-14}\\Lsun$) and redder colors than any local ULIRGs. We therefore suggest that the rise in this red fraction is predominantly an evolutionary effect due to the appearance of a more enshrouded population of luminous objects at higher redshift. The space density of this population is low; if the comoving space density remains constant with time, they may evolve into the fairly luminous local ellipticals. We speculate that we have uncovered a population of young galaxies undergoing very rapid bulge formation and / or AGN accretion, that evolve into fairly luminous galaxies ($\\sim 4L^*$) in the present-day universe. The optical-mid-IR SEDs of this population show a range of color. However, by selection, these objects are extreme in their optical-to-mid-IR properties and, as demonstrated by Figure~\\ref{color}b, have colors redder than any local dusty galaxy or ULIRG. Since we have only limited spectral information, the dust reddening cannot be determined unambiguously. However, the existing mid-infrared spectroscopy shows that these galaxies have spectra that rise into the mid-infrared, suggesting emission from warm dust, and exhibit either deep silicate absorption \\citep[e.g.,][]{hou2005,wee2006} or PAH emission \\citep[][; Desai et al. 2008, in prep]{yan2007}. Near-infrared spectroscopy by our group for a handful of galaxies reveals, where measurable, large Balmer H$\\alpha$/H$\\beta$ decrements in both the narrow and broad components of the Balmer lines \\citep[]{bra2007}. This suggests that the galaxies contain significant quantities of dust distributed on scales of, at the very least, the extended (i.e., kiloparsec scale) narrow-line emitting region. Since these galaxies lie at high redshift and are therefore very luminous, they must be undergoing significant AGN accretion and / or star-formation activity. This would suggest that the large rest-frame near-infrared to ultraviolet luminosity ratios result from heavy extinction of the most luminous emitting regions. It is difficult to estimate the extinction in these galaxies from the limited photometric information, but we can derive some general constraints by considering the extinction necessary to create such red SEDs given a young stellar population. Figure~\\ref{AV} shows the $V$-band extinction (in magnitudes) necessary to create an $(R-[24])=14$ color for template galaxies with ages less than 1~Gyr. This is based on the assumption of an extinction law derived using the functional form of \\citet[]{cal2000} for $\\lambda \\le 2.2\\mu$m and values of the extinction as tabulated in Table~4 of \\citet[]{dra2003} at longer wavelengths. Although the assumption of a monolithic dust screen is unrealistic, it does provide a lower limit on the extinction required in these galaxies. In addition, the fact that the entire galaxy is so underluminous at rest-frame UV wavelengths suggests that either the stellar population in these systems is old (i.e., $\\simgt 1$~Gyr) and intrinsically faint at UV wavelengths, or that the dust is distributed on large enough scales to shroud the stellar component as well. If these galaxies are young and dominated by a star-forming phase, then they are enveloped on kiloparsec scales by prodigious amounts of dust. Given the redshifts ($z\\sim2$) and spectral energy distributions of these galaxies, it is very unlikely that they are dominated by old stellar populations, and we therefore hypothesize that these objects are dust-obscured galaxies (DOGs), where most of the luminous regions of the galaxy are shrouded in dust. Such a scenario would suggest an overall SED for these galaxies containing three (perhaps related) components: a reddened stellar component peaking at a rest-frame wavelength of 1.6$\\mu$m, cold dust enshrouding the star-forming regions reradiating the absorbed UV emission at long wavelengths (corresponding to, say, a rest-frame wavelength of 60$\\mu$m), and warm dust enshrouding the AGN dominating the emission in the rest-frame mid-infrared (i.e., at rest-frame wavelengths of 5-20$\\mu$m). If the luminosity from the sources selected by $(R-[24])\\ge14$ is largely the result of star formation activity, then the implied star-formation rates are very large \\citep[$\\approx 1700 \\Msun$/yr for an $L_{\\rm IR}=10^{13}\\Lsun$ object;][]{ken1998b}. Such a violent starburst is unlikely to be confined to a small region, and the (observed) very large infrared-to-ultraviolet luminosity ratio must result from dust distributed on very large (kiloparsec) scales. In addition, star-formation rates this large cannot be sustained for long periods, suggesting that these galaxies are viewed during a short-lived phase during which the bulk of the stars is produced. If, instead, the sources are dominated by AGN, then the dust could be mostly confined to the nuclear regions of the galaxy. However, based on the measurement of large Balmer decrements in both broad- and narrow-line components in a few galaxies, \\citet[]{bra2007} argue that the dust distribution must be patchy and over large volumes. Even if the galaxy luminosity is dominated by AGN activity, this phase of evolution must be fairly short lived. The bolometric luminosity of an AGN may be written (in terms of its Eddington Luminosity) as $L_{\\rm bol}^{AGN}\\equiv \\epsilon L_{\\rm Edd}\\approx 0.33\\times 10^{13} (\\epsilon/0.1)(M_{\\rm BH}/10^9\\Msun) \\Lsun$, or in terms of its accretion rate as $L_{\\rm bol}^{AGN}=\\eta \\Mdot c^2 \\approx 0.15\\times 10^{13}(\\eta/0.1)(\\Mdot/1\\Msun yr^{-1})\\Lsun$. Large AGN luminosities therefore generally imply high accretion rates onto massive black holes. The present data on the DOGs do not allow any useful constraints on either the radiative efficiency, the accretion efficiency, or the black hole mass, but their luminosities of $10^{13}\\Lsun$ suggest that they must be radiating at close to the Eddington luminosity even if they harbor a very massive black hole. Even at high accretion efficiencies (i.e., $\\eta = 0.1$), the timescales to build up a $10^9\\Msun$ black hole would be short ($\\approx$0.15~Gyr). A flux density limited sample of galaxies (such as ours) will preferentially identify the galaxies in their most luminous phase, suggesting that we may be witnessing the last throes of the AGN formation in these obscured galaxies. \\begin{figure}[t] \\epsscale{1.0} \\plotone{f13.eps} \\caption{The minimum extinction required to redden the spectral energy distributions of young stellar populations and AGN to produce an optical-to-mid-infrared color of $(R-[24])=14$. The template SEDs are models from the \\citet[]{bc2003} population synthesis library with ages of 5~Myr (solid line), 25~Myr (dotted line), 100~Myr (short-dashed line) and 900~Myr (dot-dashed line), and the median QSO template (long-dashed red line) from \\citet[]{elv1994}. The reddening law used is a combination of that from \\citet[]{cal2000} and longer wavelength estimates from \\citet[]{dra2003}, and assumes the unrealistic case of a dust screen in front of the emitting source in order to derive a firm lower limit on $A_V$. \\label{AV}} \\end{figure} What is the relevance of this rather extreme population to the overall galaxy population at $z\\approx 2$? In the following subsections, we estimate the space density of the DOGs and their contribution to the infrared luminosity density in an attempt to address this question. In contrast to previous work, we do not discriminate between AGN and star-forming galaxies within our sample, but estimate their joint contribution to the space densities and contribution to the infrared luminosity density. \\subsection{Space Density of DOGs} Based on the observed redshift distribution, we now estimate the space density of this population. We assume that the comoving volume sampled by the $(R-[24])\\ge 14$ and $F_{\\rm 24\\mu m}\\ge 0.3$ selection criteria is simply the fraction of the volume between $0.5\\le z \\le 3.5$ weighted by the redshift distribution. In other words, the effective co-moving volume sampled is given by: \\begin{equation} V_c^{\\rm eff}({\\rm DOGs}) = \\Delta\\Omega \\int{g(z) dV_c} \\end{equation} where $\\Delta\\Omega$ is the solid angle covered by the survey, \\begin{equation} g(z) = {{1}\\over{\\sigma_z\\sqrt{2\\pi}}} {\\rm exp}\\left[-{{(z-\\bar{z})}\\over{2\\sigma_z^2}}\\right] \\end{equation} (with $\\bar{z}=1.98$ and $\\sigma_z=0.53$) is the (normalized) Gaussian fit to the redshift distribution, and $dV_c$ is the comoving volume element per unit solid angle per unit redshift interval \\citep[e.g.,][]{pee1993,hog1999}. For the Bo\\\"otes field, the volume sampled by the redshift distribution is roughly a third of the comoving volume between $0.5\\le z \\le 3.5$. The resulting average space density for this population in the redshift range $0.5\\le z \\le 3.5$ is \\begin{equation} \\Sigma_{\\rm DOG} (F_{\\rm 24\\mu m}\\ge 0.3~{\\rm mJy}) = (2.82\\pm 0.05)\\times 10^{-5} h_{70}^3~ {\\rm Mpc^{-3}}. \\end{equation} This space density of DOGs is comparable to that of highly UV-luminous $z\\sim 2-3$ star-forming galaxies selected using the Lyman-break or near-infrared selection techniques \\citep[e.g.,][and references therein]{red2007,vand2006}. For example, the dust-obscured galaxy population has a space density comparable to that of $M_{\\rm AB}({\\rm 1700\\AA})-5{\\rm log}h_{0.7}\\approx -22.2$ UV-bright galaxies at $z\\sim 1.9-3.4$, based on the luminosity functions derived by \\citet[]{red2007}. Although the space densities are comparable, it is important to note that each DOG contributes significantly more IR (and bolometric) luminosity per galaxy to the universe at $z\\approx 2$. The population of ``massive galaxies'' selected as ``Distant Red Galaxies'' (DRGs) at near-infrared wavelengths has a space density of $(2.2\\pm0.6)\\times 10^{-4} h_{70}^3~{\\rm galaxies~ Mpc^{-3}}$ \\citep[]{vand2006}, roughly ten times more numerous than DOGs. As mentioned above, the DRG and DOG populations also overlap in their rest-frame UV luminosity distributions, and it is therefore possible that a fraction of the star-forming DRGs are selected by our $(R-[24])\\ge 14$ criterion. A recent measurement of the luminosity function by \\citet[]{wak2006} finds that luminous (i.e., $>4L^*$) red galaxies in the local Universe have a space density of $\\approx 2.5\\times 10^{-5} h_{70}^3 {\\rm Mpc^{-3}}$, very comparable to that of the $z\\approx 2$ DOGs. It is tempting to suggest that DOGs may therefore represent a common phase of evolution for most massive galaxies. Indeed, if this phase is short lived (i.e., if the duty cycle of the enshrouded, luminous phase when galaxies may be selected as DOGs is short), then the DOGs may be the progenitors of a large fraction of present-day $>L^*$ galaxies. More detailed comparisons will require better determinations of the redshift distribution as a function of apparent magnitude and 24$\\mu$m flux density. \\subsection{Infrared Luminosity Density of DOGs} In order to estimate the infrared luminosity density contributed by the DOGs, we begin by assuming, as before, that all the DOGs with no measured spectroscopic redshifts have a redshift probability distrbution described by the Gaussian approximation to the observed redshift distribution. For the DOGs with spectroscopic redshifts, we computed their $L_8$ values and converted them to $L_{\\rm IR}$ estimates based on the procedure described above. For the DOGs without measured redshifts, we assigned random redshifts (drawn from the probability distribution) to each object and computed their luminosities. As in section \\S3.3.2, we computed luminosities using the conversion of \\citet[]{cap2007}, and derive a range based on the two extremes $L_{\\rm IR}=[5-15]\\times L_8$. Since DOGs with fainter $F_{\\rm 24\\mu m}$ are also typically fainter in the IRAC bands, we adopt the mean $\\bar{\\alpha}=2.296$ as the spectral index for all sources. The infrared luminosity density (IRLD) of the entire sample was estimated using 1000 Monte Carlo simulations of the redshift distribution of the galaxies. We find the total infrared luminosity density contributed by the DOG population to be \\begin{equation} {\\rm log[IRLD_{\\rm DOG}(F_{\\rm 24\\mu m}\\ge 0.3mJy)]} = 8.228^{+0.176}_{-0.302}, \\end{equation} where IRLD is in units of $\\Lsun~{\\rm Mpc^{-3}}$ and the limits are not 1$\\sigma$ uncertainties, but the likely range of the IRLD based on the uncertainties in the $L_8$ to $L_{\\rm IR}$ conversion (i.e., the lower and upper limits corresponding to $L_{\\rm IR}/L_8= 5$ and 15 respectively). DOGs thus provide a significant contribution to the IRLD at $z\\approx 2$. According to \\citet[]{cap2007}, the total IRLD from all 24$\\mu$m-selected $z\\sim2$ galaxies is ${\\rm log(IRLD_{\\rm Total})}\\approx 8.819^{+0.073}_{-0.079}$, with the contribution from ULIRGs alone being $42^{+15}_{-22}\\%$ of the total. Although DOGs presumably represent only a fraction of the $z\\sim 2$ population (i.e., the reddest subset), they appear to contribute $\\approx26\\pm 14\\%$ of the total IRLD from the 24$\\mu$m-selected $z\\sim 2$ galaxy population, and $\\approx 60^{+40}_{-15}\\%$ of the ${\\rm IRLD_{\\rm ULIRG}}$. The \\citet[]{cap2007} $z\\sim2$ sample is identified (in the GOODS-N and -S fields) using photometric redshift estimates and should, in principle, include DOGs. What is noteworthy, however, is that the DOGs (selected solely on the basis of extreme $R-[24]$ color) comprise such a significant component of the IRLD$_{\\rm ULIRG}$, which suggests that the DOG selection criterion effectively selects out the bulk of ULIRGs at $z\\sim 2$. Not all $z\\sim2$ galaxies may be selected as 24$\\mu$m sources. Indeed, one of the most successful methods of selecting high-redshift galaxy populations is based on the UV emission of young, star-forming galaxies \\citep[e.g.,][and references therein]{red2007}. Recently, \\citet[]{red2007} have estimated the contribution of UV-selected galaxies to the IRLD at $z\\sim2$. The IRLD contributed by DOGs is 158\\%, 60\\%, 35\\% and 56\\% of the IRLD contributed by the $L_{\\rm IR}=10^{9-10}$, $10^{10-11}$, $10^{11-12}$, and $>10^{12}\\Lsun$ UV-bright populations respectively. Do the bright and faint DOGs contribute comparably? To investigate this, we divided the DOG samples at $F_{\\rm 24\\mu m}=0.6$~mJy, resulting in 2,149 galaxies fainter than the cut with a median flux density of 0.37~mJy, and 452 galaxies brighter than the cut with a median flux density of 0.85mJy. The redshift distributions on either side of this cut are slightly different ($\\bar{z}\\approx1.9$/2.1 for the fainter/brighter galaxies with spectroscopically measured redshifts), but there are only 22 redshifts below the cut, so the distribution is not well constrained. If we assume that the distributions are the same at all flux densities, we find the IRLDs for the bright and faint subsamples are ${\\rm log[IRLD_{\\rm DOG}(F_{\\rm 24\\mu m}\\ge 0.6)]=7.78\\pm0.02}$ and ${\\rm log[IRLD_{\\rm DOG}(0.3\\le F_{\\rm 24\\mu m}< 0.6)]=7.91\\pm0.01}$ respectively. The faint and bright sources thus contribute comparable amounts to the IRLD, with the increasing number making up for the lower luminosity of the fainter source population. In actuality, there is a large range in optical-to-infrared color, and we have chosen the $(R-[24])\\ge 14$ criterion simply to isolate the higher redshift galaxies. Since the redshift distribution is broad, selecting objects at a bluer threshold would still result in high-redshift population, but with a larger low-redshift contamination. Moreover, we recall that the number density of DOGs rises toward fainter 24$\\mu$m flux densities, suggesting that $F_{\\rm 24}<0.3$mJy sources may contribute significantly to the IRLD as well. In addition, we have not corrected our IRLD estimates for the sample incompleteness at the faint 24$\\mu$m limit of our data. For all these reasons, the estimates of space density and luminosity density reported here for the DOGs are likely to be lower limits, corresponding to the contribution of only the most extreme (in $F_{\\rm 24\\mu m}$ and $(R-[24])$ color) subset of the population. It is therefore remarkable that even this relatively rare population can contribute such a significant fraction of the overall infrared luminosity density. \\subsubsection{Comparison to Other $z\\approx 2$ Galaxy Populations} The galaxies selected by virtue of their extreme mid-infrared to optical flux density ratios seem to be under-represented in most optically selected samples of high-redshift galaxies. Indeed, our 24$\\mu$m flux density criterion for DOGs (i.e., $F_{\\rm 24\\mu m}\\ge 0.3$~mJy, imposed by the existing MIPS observations of the Bo\\\"otes Field) selects a relatively rare population that has only a handful of members in most existing small-field, deep samples. In this subsection, we attempt to compare our sample with other known $z\\approx 2$ galaxy populations, most of which have been selected using observations at optical or near-infrared (i.e., rest-frame UV or optical) wavelengths. As we mentioned in \\S3.3.3 (and shown in Figures~\\ref{RbandHist} and ~\\ref{2200Hist}), the DOG population is systematically fainter than the UV-bright populations selected at $z\\approx 2$ by, say, the ``BX/BM'' selection criteria of \\citet[]{red2007}. This is not surprising, since we are selecting galaxies to have a very large mid-infrared to UV flux ratio. What is interesting is that our selection results in a sample that lies at high redshift, with no bright, low-redshift members. Moreover, it is important to note that the DOG population, while being faint at optical wavelengths, is actually bolometrically very luminous, and thereby contributes significantly to (and perhaps {\\it dominates}) the top end of the bolometric luminosity function (i.e., at $L_{\\rm IR}\\simgt 10^{12}\\Lsun$). A comparison with the $z\\approx 2-3$ ``Distant Red Galaxy\" (DRG) samples selected using near-infrared selection techniques \\citep[typically, $J-K>2.3$; e.g.,][]{franx2003,vand2004} is more difficult, since the Bo\\\"otes NDWFS field lacks very deep near-infrared imaging data. However, the DRG samples appear to have median $R_{\\rm AB}\\approx 25.9$ with 25(75)\\%-ile values of the distribution at $R_{\\rm AB}\\approx 25.1(26.7)$ \\citep[]{vand2006}. The median rest-frame luminosity of DRGs is $M_{\\rm 2200\\AA} \\approx -18.3$~AB mag for $K<21$ (Vega) samples selected from the Hubble Deep Field South and MUSYC Deep fields \\citep[G. Rudnick, personal communication;][]{gaw2006}. These magnitudes, magnitude ranges and rest-frame luminosities are comparable to those measured for our $(R-[24])\\ge14$ population, suggesting that these populations may overlap. It is now understood that the DRG selection results in a mixed sample of passively evolving systems, star-forming galaxies and AGN \\citep[e.g.,][]{kriek2006,kriek2007}. Since our selection criteria require galaxies to be bright at 24$\\mu$m, it is likely that our selection does not include the passively evolving galaxies. Recently, much work has been done using the $BzK$ selection technique pioneered by \\citet[]{dad2004} to select star-forming and passive galaxy populations at high redshift (i.e., $z\\approx 2$). We examined the $(R-24)$ colors of the $BzK$ selected star-forming galaxies in the GOODS-N field \\citep[]{dad2007a,dad2007b}, and find that only 4 of the 187 $BzK$ galaxies in GOODS-N satisfy our selection criteria. The vast majority of the $BzK$ galaxies are less extreme in their $(R-[24])$ colors, and are fainter in their 24$\\mu$m flux densities. This may be due to the lower AGN contribution (by selection) to the mid-infrared luminosities of the $BzK$ star-forming population. Given the surface density of bright DOGs (i.e., with $F_{\\rm 24\\mu m}\\ge 0.3$mJy, we would have expected roughly 14 galaxies in the GOODS-N field. Indeed, there are 21 galaxies with these selection criteria in the GOODS-N field, 3 of which are brighter than $F_{\\rm 24\\mu m}=1.0$~mJy (E. Daddi and M. Dickinson, personal communication). % Fifteen of these 21 galaxies are also selected by the $BzK$ techniques, but all but 4 are rejected from the $BzK$ `star-forming' galaxy sample because they are hard X-ray sources (E. Daddi, personal communication). The $(R-[24])\\ge 14$ and $BzK$ selections therefore appear to be comparable, with the star-forming systems at bright 24$\\mu$m flux densities effectively selected by both techniques. The simple $(R-[24])$ selection results in a slightly larger sample with perhaps a slightly broader redshift distribution. It is worth noting that the $BzK$ technique can be fairly expensive in telescope time, requiring deep imaging in three bands. The simple technique described here of using the $R$-[24] color selects similar star-forming galaxy and AGN populations more economically (90~sec with {\\it Spitzer}/MIPS and 6000~sec of $R$-band per field). Recently, a similar population of dust-obscured galaxies was identified by \\citet[]{yan2007} using the following selection criteria: $F_{\\rm 24\\mu m}>0.9$~mJy; $(\\nu F_\\nu({\\rm 24\\mu m})/\\nu F_\\nu({\\rm 8\\mu m}) > 3.16; (\\nu F_\\nu({\\rm 24\\mu m})/\\nu F_\\nu({\\rm 0.7\\mu m}) > 10$. Our selection criterion for DOGs corresponds to \\\\ $(\\nu F_\\nu({\\rm 24\\mu m}) / \\nu F_\\nu({\\rm 0.7\\mu m}) > 28.6$, and thus selects the redder, and a higher redshift subset of the population. The \\citet[]{yan2007} galaxies with $(R-[24])\\ge 14$(15) all lie at $z > 1$(1.6). The $(\\nu F_\\nu({\\rm 24\\mu m})/\\nu F_\\nu({\\rm 8\\mu m})$ distribution for DOGs shows that the selection criteria used by \\citet[]{yan2007} would reject 36\\% of the DOGs. Finally, although dust-obscured galaxies have been found in other studies using comparable selection criteria, we emphasize the simplicity and robustness of the current approach. In particular, the simple selection criteria will make it easier, in principle, to quantify the selection function once a large sample of spectroscopic redshifts become available. One significant difference in the way we select the sample of DOGs is that we do not discriminate between AGN-dominated and star-formation-dominated systems. The mid-infrared colors of many of the DOGs suggest that they may indeed be dominated by AGN at bright 24$\\mu$m flux densities (i.e., $F_{\\rm 24\\mu m}\\simgt 0.8$). Although the majority of the fainter population does show stellar ``bumps'', PAH emission features in their mid-infrared spectra, and other evidence of starlight in their spectral energy distributions, their mid-infrared, far-infrared, or bolometric luminosity may yet contain a significant contribution from an AGN. The combination of AGN and star-formation signatures in the DOG population (and in some cases in the very same objects) suggests that the population may be undergoing both rapid black-hole growth and rapid star-formation. \\subsection{The Possible Role of DOGs in Galaxy Evolution} We have demonstrated that the simple selection criterion of large mid-infrared to optical flux density ratio efficiently isolates a significant population of $z\\approx 2$ galaxies. These systems are extremely luminous at rest-frame mid-infrared wavelengths, presumably due to warm dust heated by some combination of AGN and hot, young stars. What is the role of these galaxies in galaxy evolution? In a seminal paper, \\citet[]{san1988} proposed that local ULIRGs represented a brief dust-enshrouded stage in the formation of quasars. Despite the extreme colors of the DOGs and the lack of local galaxies (even ULIRGs) with similar colors, there are some similarities between the two populations. They are both at the extreme luminous end of the galaxy population at their epochs, and they both appear to be heavily enshrouded systems exhibiting evidence for both intense star formation and AGN accretion activity. We therefore speculate that the DOGs represent a fairly short-lived phase in the evolution of massive (i.e., $\\sim4L^*$) galaxies, in between the phases represented by the sub-millimeter galaxy population (SMGs) and the more ``passive'' galaxies (e.g., selected as passive $BzK$s or $DRG$s). We envision a scenario in which gas accumulates in deep potentials, triggering an early episode of star-formation which quickly results in the formation of large quantities of dust. This results in a system which is luminous at sub-mm wavelengths (as an SMG), with cold dust temperatures. The mid-infrared spectra of these systems would show PAH features, characteristic of star-forming systems, as has been observed \\citep[e.g.,][]{men2007,pop2008}. At some point, as the accretion and star-formation proceed, an AGN is triggered which heats the dust to warmer temperatures. It is at this point that the system would have a warmer characteristic dust temperature and be selected as a DOG. If the AGN is able to destroy or expel most of the dust (or the star-formation proceeds to the point where the dust and gas are sufficiently consumed), the DOG would evolve into an optically visible AGN. Whether or not the DOG is transformed into an optically-luminous AGN will depend on the relative timescales of dust destruction / consumption and AGN accretion. It has been suggested recently that AGN feedback and star-burst driven winds can provide two important mechanisms for terminating the star-formation in galaxies \\citep[e.g.,][and references therein]{hop2007b}. In such a scenario, the galaxy evolutionary phase represented by DOGs would follow the bulge growth phase (i.e., the SMGs?), but precede the phase when the galaxy may be visible as an optical quasar or, eventually, as a red, passively-evolving galaxy (i.e., the DRGs or passive $BzK$ galaxies?). If, instead, the AGN phase is more short lived than the UV-bright star-forming phase, then DOGs would evolve into a UV-bright star-forming population. There is some circumstantial support for such a scenario from the measured space density and redshift distribution of the DOGs. The DOG surface density on the sky (of 0.089~arcmin$^{-2}$) is similar to that of the luminous sub-mm galaxy population with 850$\\mu$m flux densities $F_{\\rm 850\\mu m} > 6$mJy \\citep[]{cop2006}, and they have a comparable redshift distribution \\citep[]{cha2005}. The 850$\\mu$m sub-mm flux densities of the DOG population have not yet been measured, but the upcoming SCUBA-II Legacy survey should provide useful constraints. In the scenario described above, the DOGs are expected to have warmer dust temperatures than the SMGs, and it is indeed interesting that the detection rate of SMGs at MIPS wavelengths (24$\\mu$m and 70$\\mu$m) is low, with only $\\simlt 40$\\% of the SMGs in the GOODS-N field having 24$\\mu$m flux densities $F_{\\rm 24\\mu m}\\ge 0.3$mJy \\citep[]{pop2008}. The space density of the brightest DOGs (with $F_{\\rm 24\\mu m}\\ge 1$~mJy) is comparable to that of unobscured QSOs at comparable redshifts \\citep[e.g.,][]{bro2006}. If these brighter sources are powered primarily by AGN, they represent a significant component of the accretion history of supermassive black holes, missed by traditional QSO surveys. Clustering analyses of the population may provide a clue to the possible evolutionary products of DOGs and the lifetime of the DOG phase relative to the SMG phase, and we will investigate this in a future paper." }, "0801/0801.1051_arXiv.txt": { "abstract": "We present spectra of accretion discs around white-dwarfs calculated with an improved and updated version of the Shaviv \\& Wehrse (1991) model. The new version includes line opacities and convective energy transport and can be used to calculate spectra of hot discs in bright systems (nova--like variables or dwarf novae in outburst) as well as spectra of cold accretion discs in quiescent dwarf novae. ", "introduction": "\\label{sec:intro} In weakly magnetized cataclysmic variable stars (CVs) matter lost by a Roche-lobe filling low-mass companion forms an accretion disc around the white-dwarf primary. In bright systems such as nova-like binaries and dwarf-novae in outburst the disc is the dominant source of luminosity but even in quiescence the disc emission might provide crucial information about the physics of accretion, in particular about the mechanisms transporting angular momentum and releasing gravitational energy. Until now one could calculate only spectra for hot accretion discs \\citep[see][and references therein]{wh98}. In such an approach the disc is divided into rings whose vertical structure is calculated using the program TLUSDISK \\citet{hubeny90,hubeny91}. A different way of solving radiative transfer equations in accretion discs was presented by \\citet[][hereafter SW]{sw91}. Also, in their approach one could apply it to hot discs only. The preliminary attempt to apply the SW code to cold quiescent discs in \\citet{Idan1} was not conclusive. The main obstacle in solving the vertical structure of cold accretion discs is the presence of convection-dominated zones. \\citep[The description of the quiescent disc of SS Cyg by][neglects convection]{kromer}. The absence of spectral models for cold quiescent discs is the reason why fitting models to observations can be a frustrating exercise\\citep[see][and references therein]{urbsion}. Using the models of hot stationary discs of \\citet{wh98} to describe the spectra of cold (and non-stationary) quiescent discs cannot be expected to give relevant results. In fact since the observational data were the UV spectra obtained by the IUE, it was not very surprising that the best fits were obtained with no disc emission at all: quiescent discs are not supposed to be UV emitters so that the only source of UV radiation in such systems should be the white dwarf. The description of the quiescent state of the dwarf-nova cycle is the Achilles heel of the widely accepted disc instability model (DIM) according to dwarf-nova outbursts are due to a thermal-viscous instability \\citep[see][for a review]{l01}. For example while the DIM predicts the quiescent disc to be optically thick (especially in its outer regions), the observations of the quiescent dwarf-nova IP Peg seem to suggest the opposite \\citep{little01,ribeiro07}. Correct modeling of the emission from quiescent discs is therefore an obvious prerequisite to solving this thorny problem . In general a full description of dwarf-nova outburst cycle should include the radiation transfer equations. In practice this much too costly and therefore unfeasible. However, instead of including one can combine the DIM with the solutions of the radiative transfer equations for the same disc parameters. The \\citet[][hereafter HMDLH]{hmdl98} version of the DIM in which the time-dependent radial evolution equations use as input a pre-calculated grid of hydrostatic vertical structures (a 1+1D scheme) is well suited to such enterprize. The vertical structures are calculated using the standard equations of stellar structure or the grey-atmosphere approximation in the optically thin case. The angular-momentum viscosity transport mechanism is described by the $\\alpha$-ansatz of \\citet{ss73}. For a given $\\alpha$, $M/R^3$ and effective temperature $\\Teff$ there exist a unique solution describing the disc vertical structure. Such solutions which are very handy for solving the time-dependent equations of the disc evolution cannot be used to produce emission spectra. These can be, however, calculated by using the same input parameters ($\\alpha$, $M/R^3$ and $\\Teff$) in a radiative-transfer code on the condition that the vertical solutions calculated by the two methods are the same. In this way one can reproduce spectra of the whole cycle of a dwarf-nova outburst. The outline of the present article is as follows. In section \\ref{sec:model} we discuss the model used with special stress put on marking the differences with original SW code on which it has been based. In section \\ref{sec:results} we present solutions obtained for various types of physical set-ups. In subsection \\ref{subsec:sc} we compare our solutions for vertical structure with those obtained with HMDLH code. Then in \\ref{subsec:wh} we compare our disc spectra with the UV spectra obtained by \\citet{wh98}. In the following subsection {subsec:non-stat} we present spectra of cold non-equilibrium discs which in the DIM represent the quiescent state of the dwarf-nova outburst cycle. Section \\ref{sec:concl} contains the discussion, and conclusion. ", "conclusions": "" }, "0801/0801.4674_arXiv.txt": { "abstract": "We provide measurements of the neutral hydrogen fraction $x_{\\rm HI}$ at $z\\sim 6$, by comparing semi-analytical models of the Ly$\\alpha$ forest with observations of high-$z$ quasars and Gamma Ray Bursts absorption spectra.\\\\ We analyze the transmitted flux in a sample of 17 QSOs spectra at $5.74\\leq z_{em}\\leq 6.42$ studying separately the narrow transmission windows (peaks) and the wide dark portions (gaps) in the observed absorption spectra. By comparing the statistics of these spectral features with our models, we conclude that $x_{\\rm{HI}}$ evolves smoothly from $10^{-4.4}$ at $z=5.3$ to $10^{-4.2}$ at $z=5.6$, with a robust upper limit $x_{\\rm{HI}} < 0.36$ at $z=6.3$. We show the results of the first-ever detected transverse proximity effect in the HI Ly$\\alpha$ forest, produced by the HII region of the faint quasar RD J1148+5253 at $z=5.70$ intervening along the LOS of SDSS J1148+5251 at $z=6.42$.\\\\ Moreover, we propose a novel method to study cosmic reionization using absorption line spectra of high-redshift GRBs afterglows. We show that the time evolution and the statistics of gaps in the observed spectra represent exquisite tools to discriminate among different reionization models. By applying our methods to GRB~050904 detected at $z=6.29$, we show that the observation of this burst provides strong indications of a highly ionized intergalactic medium at $z\\sim 6$, with an estimated mean neutral hydrogen fraction $x_{\\rm HI}=6.4\\pm 0.3 \\times 10^{-5}$ along that line of sight.\\\\ \\\\ PACS 98.62.Ra - Intergalactic matter; quasar absorption systems; Lyman forest \\\\ PACS 98.54.Aj - Quasars\\\\ PACS 98.70.Rz - $\\gamma$-ray bursts\\\\ ", "introduction": "Although observations of cosmic epochs closer to the present have indisputably shown that the InterGalactic Medium (IGM) is in an ionized state, it is yet unclear when the phase transition from the neutral state to the ionized one started. Thus, the redshift of reionization, $z_{rei}$, is still very uncertain. In the last few years, our knowledge of the reionization process has been enormously increased mainly owing to the observation of $z\\sim 6$ QSOs by the SDSS survey \\cite{fan06} and CMB data \\cite{page07}. Long gamma ray bursts (GRB) may constitute a complementary way to study the reionization process possibly probing even larger redshifts, the current recold holder being GRB~050904 at $z=6.3$ \\cite{tagliaf05} \\cite{haislip06}. We provide measurements of the neutral hydrogen fraction $x_{\\rm HI}$ at epochs approaching reionization, by comparing semi-analytical models of the Ly$\\alpha$ forest with observations of the highest-$z$ QSOs and GRBs absorption spectra. ", "conclusions": "We measure the neutral hydrogen fraction $x_{\\rm HI}$ at epochs approaching the reionization, by comparing semi-analytical models of the Ly$\\alpha$ forest with observations of high-$z$ quasars and Gamma Ray Bursts absorption spectra. We consider an Early Reionization Model (ERM), characterized by a highly ionized Universe at $z\\sim 6$ and a Late Reionization Model (LRM) in which reionization occurs at $z\\sim 6$.\\\\ By comparing statistical analysis of the transmitted flux in a sample of 17 QSOs spectra at $5.74\\leq z_{em}\\leq 6.42$ with our models, we find that both ERM and LRM provide good fits to the observed LGW distribution, favoring a scenario in which $x_{\\rm HI}$ smoothly evolves from $10^{-4.4}$ at $z\\approx 5.3$ to $10^{-4.2}$ at $z\\approx 5.6$, with a robust upper limit $x_{\\rm{HI}} < 0.36$ at $z=6.3$. Discriminating among the two reionization scenarios would require a sample of QSO at even higher redshifts.\\\\ We show the results of the first-ever detected transverse proximity effect in the HI Ly$\\alpha$ forest, produced by the HII region of the faint quasar RD J1148+5253 at $z=5.70$ intervening along the LOS of SDSS J1148+5251 at $z=6.42$.\\\\ Moreover, we show that the time evolution of gaps in GRBs absorption spectra represent exquisite tools to discriminate among different reionization models. By applying our methods to GRB~050904 detected at $z=6.29$, we show that the observation of this burst provides strong indications of a highly ionized intergalactic medium at $z\\sim 6$, with an estimated mean neutral hydrogen fraction $x_{\\rm HI}=6.4\\pm 0.3 \\times 10^{-5}$ along that line of sight." }, "0801/0801.1926_arXiv.txt": { "abstract": "Mean-motion resonances (MMRs) are likely to play an important role both during and after the lifetime of a protostellar gas disk. We study the dynamical evolution and stability of planetary systems containing two giant planets on circular orbits near a 2:1 resonance and closer. We find that by having the outer planet migrate inward, the two planets can capture into either the 2:1, 5:3, or 3:2 MMR. We use direct numerical integrations of $\\sim 1000$ systems in which the planets are initially locked into one of these resonances and allowed to evolve for up to $\\sim 10^7$~yr. We find that the final eccentricity distribution in systems which ultimately become unstable gives a good fit to observed exoplanets. Next, we integrate $\\sim 500$ two-planet systems in which the outer planet is driven to continuously migrate inward, resonantly capturing the inner; the systems are evolved until either instability sets in or the planets reach the star. We find that although the 5:3 resonance rapidly becomes unstable under migration, the 2:1 and 3:2 are very stable. Thus the lack of observed exoplanets in resonances closer than 2:1, if it continues to hold up, may be a primordial signature of the planet formation process. ", "introduction": "The discovery of extrasolar planets around sun-like stars \\citep{mayorqueloz95,marcybutler95,marcybutler98} has revealed that early large-scale migration likely plays an important role in the formation of planetary systems. For example, observations have revealed a nontrivial number (roughly 6\\% overall) of close orbiting gas giants, called `hot Jupiters,' planets that would have great difficulty forming in their present locations \\citep{bodenheimeretal00}. In fact, disk-planet interaction theory predicts disturbingly short migration time scales that seem to threaten the survival of everything ranging from planetary embryos to gas giants \\citep{ward97}. Additionally, there are currently at least eight planetary systems that have two planets in MMR \\citep{udryetal07}. These include systems such as GJ 876 \\citep{marcyetal01} and HD 82943 \\citep{leeetal06} with two planets in a 2:1 resonance, and HD12661 \\citep{fischeretal03}, which contains two planets in what may be a 6:1 resonance. The origins of such resonant systems are thought to be due to disk-planet interactions that induce angular momentum transfer and differential migration \\citep{snellgroveetal01,leepeale02,papa03,kleyetal04}. There are currently two competing theories to explain planet formation. The ``core accretion\" model starts with the sedimentation and collisional growth of dust grains and smaller planetesimals in the protoplanetary disk \\citep[see][and references therein]{lissauer93}. These rocky cores (protoplanets) continue to build up until their gravitational influence allows for the accretion of the surrounding gas, forming gas giants. The other theory invokes direct gravitational instabilities in the disk \\citep[e.g.,][]{cameron78,boss00}. In these theories the dynamical relaxation time of a planetary system can, in principle, be longer than the time scale for planet formation. Therefore, long term stability is not guaranteed for a planetary system's initial configuration. Gravitational interactions between massive protoplanets can lead to orbit crossing and instability, resulting in massive planets being thrown closer to their parent stars \\citep{rasioford96,weidenschillingmarzari96}. In the case of one of the planets being ejected, the surviving planet is left with high eccentricity (some with $e > 0.90$). Additionally, migration occurs within the disk due to the exchange of angular momentum between the planet and the surrounding disk material through planet-disk interactions \\citep{goldtrem79,goldtrem80,linpapa79,linpapa93,papalin84,ward86}. In a laminar disk, there are two basic modes of migration: For an embedded body, imbalance between the torques from the inner and outer parts of the disk is thought to lead to orbital decay; this is commonly referred to as type I migration \\citep[e.g.,][]{ward97}. If a body is massive enough, it locks itself to the disk by opening a deep annular gap, and is thus carried along as the disk accretes onto the star \\citep{linpapa86,ward97}; this is called type II migration.\\footnote{There is also a proposed type III migration, which involves disk material flowing through the co-orbital region of the planet, which will not be considered here \\citep[see][]{papaetal07}.} When orbital migration occurs for different planets at different rates, convergent migration and locking into mean-motion commensurabilities can occur. For the two planet case, the regions of dynamical stability (i.e., the region where close encounters are impossible) can be determined analytically by studying how the value of the Jacobi integral relates to the topological (Hill) stability of the system \\citep{gladman93}. However, this criterion does not take into account the fact that planets may be in resonance. It has been shown that resonant systems have additional regions of stability \\citep{barnesgreen07}, which current analytic criteria predict to be unstable. In fact, \\cite{barnesgreen07} have shown that nearly all observed resonant systems lie in these extended regions. Overall, planet-planet-disk interactions are likely key in determining the final configuration of a planetary system. To date, each type of dynamical process (planet-disk interactions, dynamical instabilities, and resonances) has usually been considered individually. The motivation of this study is to explore physical situations where all of these processes occur simultaneously. In particular, we consider gap-opening planets located close to particular resonances within a protoplanetary disk. This results in convergent migration and resonance capture, but possibly also eventual instability due to close encounters. There are two objectives to this study. The first is to better quantify the regions of stability for particular resonances. More explicitly, we will consider the 3:2, 5:3, and 2:1 resonance for two gap-opening planets of varying mass. We place an upper bound on the possible masses for planets in these three resonances. This provides an interesting problem in dynamics by allowing us to more generally map the parameter space for these particular resonances in and out of the protoplanetary disk, compared to \\citet{barnesgreen07} who studied particular observed extrasolar systems. Additionally, in the case where there are instabilities, this study allows us to see how the final configurations (e.g., distribution of eccentricity) differ, if at all, from non-resonant planet scattering \\citep[see, e.g.,][]{fordetal01,chatterjeeetal07}. Our second objective is to study the dynamics of gas giants in the region of the protoplanetary disk where we believe these planets form, between 5 and 10 AU. The three above resonances are typical resonances that planets can resonantly capture into, since they require little to no eccentricity to give a high probability of capture when planets migrate at rates consistent with Type II migration. Here we are assuming that planets are born on nearly circular orbits. We then study the evolution as the coupled system of two planets migrates with the disk. To date, the closest resonance to be observed---that is, the resonance with the smallest ratio of outer to inner period---is the 2:1. We aim to better quantify why we have not found any systems in the 3:2 or 5:3 resonances and measure how common we should expect each of these resonances to be in extrasolar planetary systems. This paper is organized as follows. In \\S 2, we explain the numerical treatment used in both the resonance stability and disk simulations. In \\S 3, we give further details of the theory used in our models for studying the stability of the aforementioned resonances. We then present the results of those simulations. In \\S 4, we develop the theory used to model the planet-disk interactions in the protoplanetary disk as well as the results of those runs. We conclude with a summary and discussion in \\S 5. ", "conclusions": "In this paper, we have examined the stability of particular first and second-order resonances with and without disk interactions. We have shown that the two-planet stability criterion (computed to third-order) is invalid for these resonances (see Figures \\ref{fig:21Stab}, \\ref{fig:32Stab}, and \\ref{fig:53Stab}). The 2:1 resonance has regions of stability that extend up to mass ranges comparable to brown dwarfs, while the 3:2 resonance has a region of stability with combined planet masses up to twice the above criterion. The 5:3 resonance has a reduced region of stability compared to analytic criteria. Next, we have examined the results of instability. As shown in Figure \\ref{fig:ResCumHisto}, combining an equal number of unstable outcomes originating in each of the above three resonances (with appropriate relative fractions of ejections and collisions) yields an eccentricity distribution which fits, at least, as well as previous studies of dynamical instability in multi-planet systems (in particular \\citealt{chatterjeeetal07} for the three-planet case, and \\citealt{jurictremaine07} for larger ensembles). This is an intriguing result which raises the possibility of a link between these MMRs and exoplanet eccentricities. However, significant caution is warranted in interpreting it. Our ``outcomes\" represent snapshots at the time of instability, with no attempt made to model the subsequent effect of the disk on the remaining single planet. Also, the way in which we assemble resonances in \\S \\ref{sec:initial conditions} is rather artificial, at least in comparison to convergent migration of \\S \\ref{sec:ppdinteract}. With the addition of migration in \\S \\ref{sec:ppdinteract}, plus the eccentricity damping expected from disk interactions, we find the 2:1 resonance to be the most stable, with 100\\% of the runs remaining stable even without eccentricity damping (see Table \\ref{tab:resoutcome}). Additionally, the 3:2 resonance remains completely stable with our adopted damping prescription. A second-order resonance, the 5:3, has a high capture efficiency but goes unstable at very low eccentricities due to close encounters. This means that in order for planets to become locked into the 3:2 resonance, their primordial period separation must generally lie between 5:3 and 3:2. In this study, we selected the initial semi-major axes of the planets to coincide with the location where it is believed planets form \\citep[see, e.g.,][]{kokuboida02,thommesduncanlevison03}, rather than allowing the planets to migrate from further out in the disk. Under this assumption that planets form between 5 and 10 AU, if planets were able to form at any location with equal probability, the probability that a planet forms between the 3:2 and 5:3 resonances of a planet located at 5 AU is \\begin{equation} \\frac{\\pi\\left(\\left[(5/3)^{2/3} - (3/2)^{2/3} \\right]\\cdot 5\\right)^2}{\\pi\\cdot5^2} \\simeq 0.009. \\end{equation} That is, the probability of forming such a configuration is about 1\\%. To date, we have not discovered any planets in a closer than 2:1 resonance. The lack of 5:3 planets is readily accounted for by this study: Since this resonance is unstable at low eccentricity, it is unstable to survive all the way to a mature planetary system. The fact that 3:2 planets are {\\it also} not seen, notwithstanding the resonance's robustness, may well be telling us something fundamental about how planets form: It may simply be that neighboring giant planets are seldom born with primordial period separations of {\\it less} than 5:3. The planets in HD 12661 are believed to be in a 6:1 resonance, and there is much debate as to how they could have become locked in such a high-order resonance. This study suggests a mechanism. We have shown that an instability in one of the lower-order resonances, which needs little to no initial eccentricity for planets to become trapped in it, can result in the planets scattering apart, to then be brought together again and re-locked in a more distant, higher-order resonance (See Table \\ref{tab:rescap}). Further work is necessary to include the adjustment of the disk edge after the planets have been locked into a more distant MMR. However, migration of the outward-scattered planet would likely be somewhat more gentle in a self-consistent disk---generally we would expect an annulus of gas to initially exist between the two planets---in which case resonant re-capture would actually tend to be {\\it more} likely than in our current simple implementation. Aside from the physical findings, this study highlights an important point on orders of accuracy. Since the resonance stability tests relied on two planets of unequal mass, we chose to use up to the third order term in Equation \\ref{eqn:glad}, because this was the first term that was not symmetric in the masses (that is, a mapping $M_i \\leftrightarrow M_o$ would not be an identity mapping). Furthermore, as shown by Table 1, additional terms can change the result up to twice the first-order value. If we were to truncate Equation \\ref{eqn:glad} to just the first term, we would have concluded that the equation gives a better representation of the region of stability for the resonant case than it actually does. This paper represents a first step in our study of resonances in protoplanetary discs. Here we have constructed systems where resonance encounters were impossible to avoid: requiring that we assumed the planets were coplanar and existed near these resonances without going unstable. A natural question this paper does not address is whether the dynamics that might occur due to relative inclinations affects the overall stability of the resonant systems. It has been shown \\citep{thommeslissauer03} that it is possible for planets starting in a 2:1 resonance to later enter an inclination resonance, which generates large mutual inclinations even in an initially almost coplanar system. It is not clear \\textit{a priori} whether this would promote dynamical instability or protect against it. Additionally, we have modeled the eccentricity damping time scale as a constant proportional to the migration time scale. Recent studies \\citep{ogilvielubow03,moorheadadams07} have shown that damping may not be uniform and depends on the local properties of the disk as well as the overall geometry of the planets. In particular, $de/dt$ may even change sign depending on the value of the eccentricity. Therefore, it is important to better understand the nature of eccentricity growth and damping in order to understand the long-term stability of these resonances in the protoplanetary disk. Our improved distribution shown in Figure \\ref{fig:ResCumHisto} was derived by using an equal number of cases from each of the three resonances studied. However, Table \\ref{tab:resoutcome} suggests that we should find more planets in the 2:1 resonance compared to the 3:2 resonance during the disk's lifetime when we assume that all the planets migrated from a region farther out than the 2:1 resonance. In the case of gravitational instability, it may be assumed that planets could form in a more equally distributed manner before undergoing migration and locking into these resonances. Complete three dimensional integrations would allow us to gain an accurate representation of the distribution of resonance captures due to migration alone. Further stability tests will involve allowing one planets to migrate from greater distances, where near-resonance interactions cannot affect the overall stability of the system until the planets (given they remain stable) can migrate close enough together to encounter a particular resonance. Additionally, gap-forming planets on the edge of a disk in general ought to still be accreting some mass across their gap, and therefore mass growth will be taken into consideration. In this case, planets could lock into resonance under conditions we have shown to be stable, and through mass accretion enter a regime that analytic and numerical criteria determine to be unstable. A comparison of whether this region of stability is similar to what we have found here for constant-mass planets would give us further important insights into the possible range of planetary system configurations. We have gained a clearer picture of the range of possible dynamics within a protoplanetary disk. In doing so, we have seen how resonances can strongly affect planetary systems, long before the surrounding disk material has dissipated. This suggests that what happens before the disk disappears is key in determining a planetary system's ultimate dynamical architecture. One possible outcome is a system which, after the disk is gone, is left with planets in resonance. In order for this to happen, the planets must enter these resonances relatively late in the disk's lifetime, so that they ultimately survive the coupled migration. After the disk has dissipated, some resonances may eventually become unstable. Another possibility we have demonstrated is that resonances can be broken by dynamical instability that occurs while the disk is still present. Our results allow us to make a connection between the planet formation process and observations of mature, resonant planetary systems, none of which have yet been observed in a closer than 2:1 resonance: For planets that form further apart than a period ratio of 2:1 and migrate convergently, the very stable 2:1 constitutes a formidable barrier. Between 2:1 and 5:3, the initial outcome is likely to be capture into 5:3, followed by a one-way trip to instability. The 3:2 resonance is quite stable (unless eccentricity damping by the disk is absent), so provided that an appreciable number of planets form with a primordial period separation of less than the 5:3, we would expect to find some planets in a 3:2 resonance. Thus, if future observations continue to reveal no such systems, then this may suggest a lower limit to orbital separations with which neighboring planets first form." }, "0801/0801.4732_arXiv.txt": { "abstract": "{This is the third of a series of papers presenting the first attempt to analyze the growth of the bar instability in a consistent cosmological scenario. In the previous two articles we explored the role of the cosmology on stellar disks, and the impact of the gaseous component on a disk embedded in a cosmological dark matter halo.} {The aim of this paper is to point out the impact of the star formation on the bar instability inside disks having different gas fractions.} {We perform cosmological simulations of the same disk-to-halo mass systems as in the previous works where the star formation was not triggered. We compare the results of the new simulations with the previous ones to investigate the effect of the star formation by analysing the morphology of the stellar components, the bar strength, the behaviour of the pattern speed. We follow the gas and the central mass concentration during the evolution and their impact on the bar strength.} {In all our cosmological simulations a stellar bar, lasting 10 Gyr, is still living at z=0.\\\\ The central mass concentration of gas and of the new stars has a mild action on the ellipticity of the bar but is not able to destroy it; at z=0 the stellar bar strength is enhanced by the star formation. The bar pattern speed is decreasing with the disk evolution.} {} ", "introduction": "The connection between the bar feature and the star formation process has already been pointed out in the past: \\citet{mart77} observed that non interacting galaxies displaying the highest star forming activity have strong and long bars. Conversely not all strong and long bars are intensively creating stars. \\citet{Ma01} performed SPH simulations to investigate the dependence of the star formation in disks on the geometry and dynamical state of the DM halo. They showed that the star formation lengthens the life time of the bar. However in their works the evolution of the disk+halo system arises in a isolated framework, outside the cosmological scenario.\\\\ The treatment of the star formation in galaxies involves a large number of physical processes which arise with different time scales and spatial lengths. Therefore its inclusions in simulations of galaxy formation is complex (\\citet{MaCu03} and references therein) and many sub-grid parametrisations (i.e. representing all the astrophysical processes that cannot be resolved due to resolution or computing power limitations) of the star formation process, exist in literature. To mimic all these processes with a numerical model of a realistic disk galaxy is difficult too. Thus the numerical work on this subject focuses either on a detailed analysis of the interstellar medium (ISM) and a smaller simulated area \\citep{Wa01, avil2000} or on the study of the global instabilities and of the star formation at the cost of simplified ISM models \\citep{Spri04, Sem02, Comb02}. Models that have tackled both topics have either been in two dimensions \\citep{WaNo01} or restricted to a box size a few hundred parsecs across \\citep{Wa01}. \\citet{Ros95} worked in two dimensions allowing the ISM to evolve self-consistently but treated the stars as collisional rather than collisionless fluid. \\citet{task06} presented the first three dimensional simulations of a global disk without the need to simplify the structure of the ISM. Their paper is devoted to understand the fundamental processes of star formation and feedback in a disk galaxy.\\\\ Here we performed cosmological simulations to analyse the effects of the star formation process inside stellar-gaseous disks embedded in a DM cosmological halo evolving in a fully consistent scenario. We employ the sub-grid star formation model by \\citet{Spri03} which is able to produce a self-regulated star formation in galactic disks. We point out that our model is not a general galaxy evolution model, since the gradual formation and growth of the stellar disk has not been taken into account. However our approach, developed in two previous papers \\citep{Cu06, Cu07}, allows us to investigate the effects of different parameters like the disk-to-halo mass ratio and the gas fraction inside the disk, on the growth of the bar instability, its coupling with the star formation rate and its dependence on such parameters in a self-consistent cosmological framework. We compare the results of such a new set of simulations with those of our previous sets with the star formation switched off (\\citet{Cu06}, hereafter Paper 1, \\citet{Cu07}, hereafter Paper 2). In Paper 1 we presented simulations of purely stellar disks with the same disk-to-halo mass ratios, in the same cosmological scenario, and with the same initial conditions as in this new paper. In Paper 2 we investigated the growth and the evolution of the bar instability in disks with the same disk-to-halo mass ratios as in Paper 1, and different gas fractions, without star formation. Thus, we will compare the results of simulations here performed, with those corresponding to the same disk-to-halo mass ratio as in Paper 1 and to the same gas fraction as in Paper 2. The comparison between the parameters characterising the bar can be done only evaluating that of the old stars component, since the new stars are not present in Paper 1 and 2. In both such Papers it was shown that in DM dominated disks a bar feature is triggered and maintained by the Cosmology whereas, in the more massive disks, a gas fraction 0.2 is able to destroy the bar. The focus of the present work is to determine if the star formation changes such a result. The plan of the paper is the following: in Section 2, we summarise our recipe for the initial $disk+halo$ system fully described in Paper 1, and present our star formation recipe. In Section 3 there are our cosmological simulations, in Section 4 we point out our results. The parameters related to the bars formed in the new and the old stars and to the global, old+new, stellar populations are given in such a section. Section 5 is devoted to our discussion and conclusions. It contains a Table including the parameters characterising the old stellar bar at z$=0$ in the three Papers to allow a more easy comparison between the results of our Papers (Table 2). ", "conclusions": "We presented six cosmological simulations with the same disk--to--halo mass ratios as in Paper 1 and Paper 2. In order to study the impact of the forming stars from the gaseous component, here we included and varied its percentages inside disks of different disk-to-halo mass ratios, as in Paper 2 where the star formation was switched off. In Table \\ref{fin_table} we show the final data for ellipticities and semimajor axes resuming the results of our three papers. The old star component show a long lasting bar, 10 Gyr old, in all the simulations of this work. \\begin{table*} \\caption{Cosmological disks simulations: global results} \\label{fin_table} \\centering \\begin{tabular}{c c c c c c } \\hline\\hline Disk mass& gas fraction & {$ $} no star formation & {}& star formation & {}\\\\ {}& {} & ellipticity& $a_{max}$ & ellipticity & $a_{max}$ \\\\ \\hline 0.33 & 0. & 0.52 & 5 & - & - \\\\ 0.33& 0.1 & 0.68 & 8.4 & 0.65 & 8 \\\\ 0.33& 0.2 & 0.1 & no bar & 0.55 & 11 \\\\ 0.33& 0.4 & 0.07 & no bar & 0.6 & 8.4 \\\\ 0.1& 0. & 0.3 & 6.5 & - & - \\\\ 0.1&0.1 & 0.58 & 5.8 & 0.39 & 3 \\\\ 0.1& 0.2 & 0.6 & 5.4 & 0.45 & 3 \\\\ 0.1& 0.6 & 0.42 & 5.8 & 0.5 & 3 \\\\ \\hline \\end{tabular} \\end{table*} Moreover, in all such simulations, except for simulation c6, the bar is stronger than that developed in the pure stellar case with the same disk-to-halo mass ratio but weaker than that formed in the case of the same gas fraction without star formation. We find that the star formation, reducing the central gaseous mass concentration, allows the bar to survive until the end of the evolution also in the more massive disks, at variance with the results in Paper 2 (where the gas was not allowed to form stars): in such a case a gas fraction 0.2 was able to destroy the bar. Even if some details of the disk morphologies obviously depend on the star formation prescription, the bars arising here in the c2 and c3 simulations are due to the reduced central gas concentration and thus to the weakening of its effect on the bar itself. We verified this point by rerunning simulation c2 changing the star formation prescription. Instead of the GADGET-2 effective model, we used a simple algorithm in which, when a gas particle reaches a given density threshold with a temperature lower than a minimum temperature threshold, it is converted into a star particle, as in \\citet{Katz96}. Fig. \\ref{katz} shows the morphology of the old stellar component for such a different star formation recipe: the survival of the bar is due to the star formation, independently on its details.\\\\ In all the simulations of more massive disks, the new stellar component shows a barred shape coupled with that of old stars but with semimajor axis and ellipticity values smaller than those. \\begin{figure} \\begin{center} \\includegraphics[width= 7cm]{isoold_xyz_levfixG179_12lev_rot35_Katz.ps} \\end{center} \\caption{ Isodensity contours of the old stars at z=0, for the simulation c2, rerunned with the Katz'96 star formation recipe} \\label{katz} \\end{figure} In all the simulations of DM dominated disks, a bar feature is maintained in the old stellar component at the disk center, and the new stars form a spheroidal bulge that hides the bar of the old stars. This effect could be of interest as far as the observations about the Milky Way are concerned where, in addition to the main bar, a small nuclear bar consisting of old stars seems to be included in the bulge (see e. g. \\citet{ala01}). The final strenght of these old star bars increases by increasing gas fraction and their pattern speed is quickly decreasing before z$=1$. The classical results obtained outside the cosmological scenario are no longer applicable. This conclusion remarks the results of Paper 1, where it was shown that in the DM dominated disks the bar feature is triggered and maintained by the cosmological properties of the halo (namely its triaxiality and its dynamical state). From the models presented here we suggest that a very low pattern speed (few $ Km\\, s^{-1}\\, Kpc^{-1}$) could be a signature of a dominating halo and not a classical product of the disk instability. The whole set of cosmological simulations we presented here is suggesting that the star formation works in favour of maintaining the bar feature in self gravitating disks, against the effect of increasing gas fractions, whereas in DM dominated disks, it slightly reduces the bar strength. Moreover, inside such disks, the new stars are arranged in non barred systems which are decoupled from the bar of the old star component. {\\bf Acknowledgements} Simulations have been performed on the CINECA IBM CLX Cluster, thanks to the INAF-CINECA grants cnato43a/inato003 ``Evolution of disks in cosmological contests: effect of the star formation inside the disk'', and on the Linux PC Cluster of the Osservatorio Astronomico di Torino. We wish to thank V. Springel for kindly providing us with his code GADGET, and Martina Giovalli which implemented the simple star formation prescription in it. We thank the referee for his suggestions, useful to improve the paper." }, "0801/0801.2578_arXiv.txt": { "abstract": "The frequency of microlensing planet detections, particularly in difficult-to-model high-magnification events, is increasing. Their analysis can require tens of thousands of processor hours or more, primarily because of the high density and high precision of measurements whose modeling requires time-consuming finite-source calculations. I show that a large fraction of these measurements, those that lie at least one source diameter from a caustic or the extension from a cusp, can be modeled using a very simple hexadecapole approximation, which is one to several orders of magnitude faster than full-fledged finite-source calculations. Moreover, by restricting the regions that actually require finite-source calculations to a few isolated ``caustic features'', the hexadecapole approximation will, for the first time, permit the powerful ``magnification-map'' approach to be applied to events for which the planet's orbital motion is important. ", "introduction": "\\label{sec:intro}} Microlensing planets are being discovered at an accelerating rate, with one being reported in 2003 \\citep{ob03235}, three in 2005 \\citep{ob05071,ob05390,ob05169}, two in 2006 \\citep{gaudi08}, and perhaps as many as six in 2007. It has been a huge challenge for modelers to keep up with these discoveries, in large part because the computing requirements are often daunting: the parameter space is large, the $\\chi^2$ surface is complex and generally contains multiple minima, and the magnification calculations are computationally intensive. Proper treatment of individual events can require tens of thousands processor hours, or more. Indeed, some potential planetary events have still not been fully modeled because of computational challenges. Planetary microlensing events are recognized through two broad channels, one in which the lightcurve perturbation is generated by the so-called ``planetary caustic'' that is directly associated with the planet \\citep{gouldloeb92} and the other in which it is generated by the ``central caustic'' that is associated with the host star \\citep{griestsafi}. The former events are relatively easy to analyze and indeed the event parameters can be estimated reasonably well by inspection of the lightcurve. The latter events are generally much more difficult. There are several interrelated reasons for this. First, planets anywhere in the system can perturb the central caustic. This is why these events are avidly monitored, but by the same token it is often not obvious without an exhaustive search which planetary geometry or geometries are responsible for the perturbation. Second, this very fact implies that several members of a multi-planet system can be detected in central-caustic events \\citep{gaudi98}. Multiple planets create a larger, more complicated parameter space, which can increase the computation time by a large factor. Even if there are no obvious perturbations caused by a second planet, an exhaustive search should be conducted to at least place upper limits on their presence. Third, if the source probes the central caustic, it is ipso facto highly magnified. Such events are brighter and more intensively monitored than typical events and so have more and higher-quality data. While such excellent data are of course a boon to planet searches, they also require more and more-accurate computations, which requires more computing time. Fourth, central-caustic planetary events are basically detectable in proportion to the size of the caustic, which roughly scales $\\propto q/|b-1|$, where $q$ is the planet/star mass ratio and $b$ is the planet-star separation in units of the Einstein ring. Thus, these events are heavily biased toward planets with characteristics that make the caustic big. Such big caustics can undergo subtle changes as the planet orbits its host, and in principle these effects can be measured, thus constraining the planet's properties. Exploration of these subtle variations requires substantial additional computing time. Finally, there is also a bias toward long events, simply because these unfold more slowly and so increase the chance that they will be recognized in time to monitor them intensively over the peak. Such long events often display lightcurve distortions in their wings due to the Earth's orbital motion which, if measured, can further constrain the planet properties. However, probing this effect (called ``microlens parallax'') requires yet another expansion of parameter space. Moreover, the ``parallax'' signal must be distinguished from ``xallarap'' (effects of the source orbiting a companion), whose description requires a yet larger expansion of parameter space. There are two broad classes of binary-lens (or triple-lens) magnification calculations: point-source and finite-source. The former can be used whenever the magnification is essentially constant (or more precisely, well-characterized by a linear gradient) over face of the source, while the latter must be used when this condition fails. Point-source magnifications can be derived from the solution of a 5th (or 10th) order complex polynomial equation and are computationally very fast. When the point-source is well separated from the caustics (as it must be to satisfy the linear-gradient condition) then this calculation is also extremely robust and accurate. The main computational challenge in modeling planetary events comes from the finite-source calculations. Almost all integration schemes use inverse-ray shooting, which avoids all the pathologies of the caustics by performing an integration over the {\\it image} plane (where the rays behave smoothly) and simply asks which of the rays land on the source. The problem is that a large number of rays must be ``shot'' to obtain an accurate estimate of the magnification, which implies that high-quality data demand proportionately longer computations. Of course, the higher the magnification, the larger the images, and so the more rays are required. There are various schemes to expedite inverse ray-shooting, including clever algorithms for identifying the regions that must be ``shot'' and pixelation of the source plane. The bottom line is, however, that the overwhelming majority of computation time is spent on finite-source calculations. Here I present a third class of binary-lens (or triple-lens) computation that is intermediate between the two classes just described: the hexadecapole approximation. \\citet{pejcha07} expand finite-source magnification to hexadecapole order and illustrate that the quadrupole term by itself can give quite satisfactory numerical results. In this paper, I develop a simple prescription for evaluating this expansion. While this algorithm is about 10 times slower than point-source calculations, it is one to several orders of magnitude faster than finite-source calculations. The method can be applied whenever the source center is at least two source radii from a caustic or the extension from a cusp. Typically, well over half the non-point-source points satisfy this condition, meaning that the method can reduce computation times by a factor of several. Moreover, by isolating the small regions of the lightcurve where finite-source calculations must be used, the method opens up the possibility that the finite-source computations themselves can be radically expedited for the special, but very interesting, class of events in which planetary orbital motion is measured. ", "conclusions": "\\label{sec:conclude}} I have identified an intermediate regime between the ones where finite-source effects are dominant and negligible, respectively. In this regime, magnifications can be evaluated with very high precision using a simple hexadecapole approximation, for which I give a specific prescription. Outside of small (few source-diameter crossing times) regions associated with caustic crossings and cusp approaches, the approximation has a fractional error of well under 0.1\\%. Some events can now be analyzed without any traditional finite-source calculations, while for others these calculations will be drastically reduced. In particular, by restricting the regions that do absolutely require finite-source calculation to a few isolated zones, this approximation opens the possibility of applying ``magnification maps'' to planetary systems experiencing significant orbital motion." }, "0801/0801.4865_arXiv.txt": { "abstract": "We construct self-consistent dynamical models for disk galaxies with triaxial, cuspy halos. We begin with an equilibrium, axisymmetric, disk-bulge-halo system and apply an artificial acceleration to the halo particles. By design, this acceleration conserves energy and thereby preserving the system's differential energy distribution even as its phase space distribution function is altered. The halo becomes triaxial but its spherically-averaged density profile remains largely unchanged. The final system is in equilibrium, to a very good approximation, so long as the halo's shape changes adiabatically. The disk and bulge are ``live'' while the halo is being deformed; they respond to the changing gravitational potential but also influence the deformation of the halo. We test the hypothesis that halo triaxiality can explain the rotation curves of low surface brightness galaxies by modelling the galaxy F568-3. ", "introduction": "Dark matter halos -- at least the ones found in cosmological simulations -- have a number of universal traits. Most famously, their density profiles have a shape that is nearly independent of mass, formation epoch, and cosmological model \\citep{nfw96}. Their angular momentum distribution \\citep{bullock01}, phase space density \\citep{taylor01}, and velocity anisotropy \\citep{hansen06} profiles also appear to follow universal forms. In other respects, halos are rather diverse. Simulated halos are typically triaxial with axis ratios that range from $0.6$ to $1$ (See \\citet{dubinski91, warren92} and more recently, \\citet{novak06}). The shapes of real halos are more difficult to determine but promising observational approaches do exist. Probes of the Galactic halo include flaring of the gas disk \\citep{olling} and tidal streams of satellite galaxies \\citep{johnston99}. The shapes of halos in other galaxies can be determined, at least statistically, by weak gravitational lensing surveys \\citep{hoekstra, mandelbaum, parker}. On the other hand, triaxiality can bias attempts to determine a halo's density profile from the rotation curve of the disk that sits within it \\citep{hayashi06}. Our main goal in writing this paper is to introduce a novel scheme to generate self-consistent dynamical models for disk galaxies with triaxial halos. Our models can be tailored to fit observational data for specific galaxies and therefore provide a testing ground to study the disk-halo connection. We consider the effects of halo triaxiality on the rotation curves of low surface brightness galaxies (LSBs) and briefly discuss other applications of the models. Halos in simulations of a cold dark matter (CDM) universe have central cusps with $\\rho\\propto r^{-\\gamma}$ where $\\gamma\\simeq 1$ \\citep{nfw96}. In dark matter-dominated galaxies, this density profile would seem to imply a rotation curve where $v\\propto r^{1/2}$ as $r\\to 0$. By contrast, a halo with a constant density core implies $v\\propto r$ as $r\\to 0$. LSBs, which are believed to be dark-matter dominated at small radii, have rotation curves that generally favor a constant density core over a $\\gamma=-1$ cusp. This result represents one of the most serious challenges to the CDM scenario \\citep{moore94, flores94, mcg98} and has inspired some rather exotic alternatives. De Blok \\& McGaugh (1998), for example, suggested that LSB rotation curves could be explained by Modified Newtonian Gravity while \\citet{firmani00} and \\citet{mo00}) invoked dark matter self-interactions to flatten the central cusp of the halo. The connection described above between a galaxy's rotation curve and the intrinsic density profile of its halo assumes that the halo is spherically symmetric. However, if a galaxy's halo is triaxial, then gas in the disk will move on non-circular orbits and, under certain conditions, the observed rotation curve will rise approximately linearly even if the intrinsic halo density profile has a steep cusp. \\citet{hayashi06} and \\citet{hayashi07} presented this argument as a means of reconciling LSB rotation curves with the predictions of the CDM model. \\citet{hayashi06} and \\citet{hayashi07} derived model rotation curves by calculating closed orbits in the potential generated by a triaxial halo. In this paper, we derive rotation curves by making pseudo-observations of a disk that is embedded in the halo. Deviations from axial symmetry in disk and halo are generated concurrently and self-consistently. A number of methods exist for constructing models of, and embedding disks in, triaxial halos. For example, \\citet{moore04} show that the remnant of a major merger between two equilibrium spherical halos is triaxial. \\citet{bailin07} describe how to set up an equilibrium disk in a combined halo-disk potential. Our approach produces an N-body galaxy complete with disk, bulge, and triaxial halo. (Central black holes may also be included, as in \\citet{wid05}.) It is inspired by the method outlined in \\citet{holley01}. In that scheme, dubbed ``adiabatic squeezing'', particles of an equilibrium halo are subjected to an artificial drag by modifying the force law of a standard N-body code and evolving the system forward in time. A triaxial halo is created if the drag has a different strength along three orthogonal directions, that is, along what become the three principle axes of the halo. The final model will be in equilibrium, to a good approximation, so long as the timescale for the halo's shape to change is slow as compared to the typical orbital timescale of the system. Adiabatic squeezing causes a halo to shrink in size. For an isolated halo, this shrinking can be reversed by simply rescaling the positions and velocities of the particles. Obviously, this method is unsuitable for disk-bulge-halo systems since the disk and bulge would be disrupted in an unphysical way. We propose a modification of this method in which drag is applied along one axis and ``negative drag'' is applied along the other two or vice versa depending on whether one wants a prolate or oblate halo. We require that for each particle, the change in energy due to the artificial drag force is zero. In this way, we change the phase space distribution of the particles but not their energy distribution. As noted in \\citet{BT}, {\\it if two systems have the same energy distribution, their spherically-averaged density profiles will be very similar even if their phase space distribution functions are different.} Our starting point is the equilibrium model of \\citet{wid07} which comprises a Sersic bulge, cuspy dark halo, and exponential disk. The model is described in terms of a phase space distribution function (DF) which, in turn, is a function of the integrals of motion. In the current version of the model, the halo component of the DF depends only on the energy. In the absence of a disk, the halo is spherically symmetric. With the disk included, the halo is flattened slightly but is still axisymmetric. Adiabatic deformation allows us to extend our disk-bulge-halo model to systems with triaxial halos. In Section 2, we describe our method and construct an example of an isolated triaxial halo. We then consider the LSB galaxy F568-3. In Section 3, we present axisymmetric, equilibrium models for this galaxy based on its published surface brightness profile and circular speed curve. In Section 4, we show how transforming the axisymmetric halo in one of these models into a triaxial halo changes the shape of the rotation curve. We conclude in Section 5, by summarizing our results and briefly discussing further applications of the method. ", "conclusions": "The adiabatic squeezing method \\citet{holley01} produces triaxial halos that have shrunk in size and therefore requires that the positions and velocities of the particles be rescaled. This awkward step precludes the technique from being applied to compound systems. Our approach avoids this problem by using an {\\it energy-conserving} artificial force to deform the halos. Our analysis of the LSB galaxy F568-3 begins with a discussion of axisymmetric models. We attempt to fit both photometric and kinematic observations using Bayesian statistics and the MCMC method. Our excellent fit of the surface brightness profile requires a bulge and disk truncation, neither of which were included in previous studies. As for the rotation curve, we find that constant density cores do better than density cusps in agreement with earlier studies of LSBs. The second stage of our analysis is to deform the halo of a compound system. In agreement with \\citet{hayashi06}, we show that the rotation curve of F568-3 may indicate the presence of a triaxial halo rather than a problem with the standard CDM model of structure formation. \\citet{hayashi06} and \\citet{hayashi07} construct rotation curves by finding closed orbits in the gravitational potential of a triaxial halo. We calculate the rotation curves by making pseudo-observations of a disk that is self-consistently embedded in a dark halo. There are two improvements that will add a further level of realism to the analysis: the inclusion of a gas disk in the galactic models and an iterative scheme whereby the initial model and artificial acceleration parameters are adjusted so that the final model fits the data in detail. These improvements will be considered in a future publication. Our triaxial models have a wide range of applications. For example, they can be used to study the effect a non-spherical halo has on the morphology of tidal streams from satellite galaxies and flaring and warping of the gas disk. The method can also be applied to bulges where departures from axial symmetry are thought to be important." }, "0801/0801.3959_arXiv.txt": { "abstract": "We demonstrate the ability to measure precise stellar barycentric radial velocities with the dispersed fixed-delay interferometer technique using the Exoplanet Tracker (ET), an instrument primarily designed for precision differential Doppler velocity measurements using this technique. Our barycentric radial velocities, derived from observations taken at the KPNO 2.1 meter telescope, differ from those of Nidever et al. by 0.047 km s$^{-1}$ (rms) when simultaneous iodine calibration is used, and by 0.120 km s$^{-1}$ (rms) without simultaneous iodine calibration. Our results effectively show that a Michelson interferometer coupled to a spectrograph allows precise measurements of barycentric radial velocities even at a modest spectral resolution of R $\\sim 5100$. A multi-object version of the ET instrument capable of observing $\\sim$500 stars per night is being used at the Sloan 2.5 m telescope at Apache Point Observatory for the Multi-object APO Radial Velocity Exoplanet Large-area Survey (MARVELS), a wide-field radial velocity survey for extrasolar planets around TYCHO-2 stars in the magnitude range $7.62.5$, and orange circles are galaxies with $n<2.5$. The dashed lines show the least-squares linear fit to the early-type galaxies, i.e., the magenta points in the top panel, iteratively rejecting $3\\sigma$ outliers. The solid lines are the same as the dashed lines but shifted blue-ward by $2\\sigma$.} \\label{M_gr} \\end{figure} ", "conclusions": "We have shown that structure (in this Letter quantified by the S\\'ersic parameter $n$) and morphology (quantified by $n$ and the bumpiness parameter $B$) are distinct galaxy properties. There is a significant number of galaxies with high S\\'ersic indexes but late-type morphologies (Fig. \\ref{n_col}). The physical significance of the difference between structure and morphology becomes apparent when their behavior as a function of galaxy mass and environment is analyzed. Structure mainly depends on galaxy mass whereas morphology, at fixed galaxy mass, depends on environment (Fig. \\ref{struct}). The different behavior is linked to star formation activity, which only weakly affects the structure of a galaxy, but strongly affects its morphological appearance (Fig. \\ref{colorB}). This implies that the existence of the MDR is intrinsic, and is explained by the environmental dependence of star formation activity. This seems trivial; however, it means that the MDR is not simply the result of underlying correlations, i.e., the strong dependence of structure and star formation on galaxy mass, and the environmental dependence of the mass function. The author gratefully acknowledges financial support from NASA grant NAG5-7697, and is indebted to support from John Blakeslee and Holland Ford. Stimulating discussions with Andrew Zirm and Marijn Franx helped shape this Letter. Attentive reading by and valuable comments from Andrew Zirm, Roderik Overzier, Alison Crocker, John Blakeslee, Bradford Holden, Ricardo Demarco, and Marc Postman have greatly improved the readability." }, "0801/0801.3397_arXiv.txt": { "abstract": "Very massive primordial stars ($140\\ M_{\\odot} < M < 260\\ M_{\\odot}$) are supposed to end their lives as PISN. Such an event can be traced by a typical chemical signature in low metallicity stars, but at the present time, this signature is lacking in the extremely metal-poor stars we are able to observe. Does it mean that those very massive objects were not formed, contrarily to the primordial star formation scenarios ? Could it be possible that they avoided this tragical fate ? We explore the effects of rotation, anisotropical mass loss and magnetic field on the core size of very massive Population III models. We find that magnetic fields provide the strong coupling that is lacking in standard evolution metal-free models and our $150\\ M_{\\odot}$ Population III model avoids indeed the pair-instability explosion. ", "introduction": "According to \\citet{HFWLH03}, the fate of single stars depends on their He-core mass ($M_{\\alpha}$) at the end of the evolution. They have shown that at very low metallicity, the stars having $64\\ M_{\\odot} < M_{\\alpha} < 133\\ M_{\\odot}$ will undergo pair-instability and be entirely disrupted by the subsequent supernova. This mass range in $M_{\\alpha}$ has been related to the initial mass the star must have on the main sequence (MS) through standard evolution models: $140\\ M_{\\odot} < M_{\\rm ini} < 260\\ M_{\\odot}$. Since we will present here a non-standard evolution, we will rather keep in mind the $M_{\\alpha}$ range. ", "conclusions": "The model we presented here is exploratory. We cannot draw general conclusions from it. But our model shows that heavy mass loss is possible even at $Z=0$, and the answer to our title's question is: \\emph{yes, under certain conditions, very massive stars can indeed avoid PISN}. Some aspects need yet to be clarified. Is the WR mass loss rate we have used really valid? Can the CNO lines alone really drive a WR wind? In a more general perspective, we still lack a good mass loss recipe for the strict $Z=0$ case. A word of caution must also be cast on the inclusion of magnetic fields. The validity of the Tayler-Spruit dynamo is still under debate, and more work need to be done before we may confidently rely on results that have been obtained with the actual treatment. Moreover, there is not a clear consensus whether magnetic fields were present in the early Universe or not. In the Taylor-Spruit dynamo, the mechanism amplifies a pre-existing field, so if none were present at the start, we cannot use it. Anyway, the physics used in the present model is today's ``state of the art'' and it is interesting to study what can be achieved with it. Our result is encouraging, because the computation has been accomplished with reasonable assumptions: the initial rotation rate used here was fast but not extreme, and the mechanisms called upon (anisotropy of the winds and magnetic fields) are not exotic ones, but two natural effects which arise when one treats properly the case of rotation. After this first step, more work is needed. First, we have to refine the mass loss treatment at $\\Omega\\Gamma$-limit and check if the mass loss stays as strong as here. If it does so, we have to check if our result are valid in the whole PISN mass domain. Higher mass models should experience higher mass loss, but it is necessary to check if it would still be sufficient to help them avoid pair-instabilities." }, "0801/0801.3674_arXiv.txt": { "abstract": "We study the spectrum of cosmological fluctuations in the D3/D7 brane inflationary universe with particular attention to the parametric excitation of entropy modes during the reheating stage. The same tachyonic instability which renders reheating in this model very rapid leads to an exponential growth of entropy fluctuations during the preheating stage which in turn may induce a large contribution to the large-scale curvature fluctuations. We take into account the effects of long wavelength quantum fluctuations in the matter fields. As part of this work, we perform an analytical analysis of the reheating process. We find that the initial stage of preheating proceeds by the tachyonic instability channel. An upper bound on the time it takes for the energy initially stored in the inflaton field to convert into fluctuations is obtained by neglecting the local fluctuations produced during the period of tachyonic decay and analyzing the decay of the residual homogeneous field oscillations, which proceeds by parametric resonance. We show that in spite of the fact that the resonance is of narrow-band type, it is sufficiently efficient to rapidly convert most of the energy of the background fields into matter fluctuations. ", "introduction": "In recent years a lot of interest has focused on the interface area between superstring theory and cosmology. The reasons are twofold. Firstly, cosmology can provide a possible arena to test string theory observationally. Secondly, superstring theory may be able to resolve the conceptual problems from which the current realizations of scalar field-driven inflation suffer (see e.g. \\cite{RHBrev} for a discussion). In particular, a lot of attention has recently been devoted to attempts to obtain periods of inflationary expansion of space in compactifications of superstring theories to four space-time dimensions. Since such compactifications contain a large number of scalar fields and since some of those remain massless before supersymmetry breaking, string theory gives rise to promising candidates for the inflaton, the scalar field driving the inflationary expansion of space (see e.g. \\cite{Linderev,Cliffrev,JCrev,EvaLiam} for recent reviews of attempts to obtain inflation from string theory compactifications). D-branes have played a particularly important role in the recent approaches to obtaining inflation from string theory \\cite{Dvali,Stephon,Shafi,Cliff,JuanGB,KKLMMT}. A particularly promising model is the D3/D7 brane inflation model, in which the inflaton is the separation between the two branes in the directions transverse to both branes \\cite{DHHK}. This model features a shift symmetry \\cite{shift} for the inflaton field which ensures that at the classical level, the potential is flat along the inflaton direction (this shift symmetry is broken under quantum corrections \\cite{kors}). Inflation is induced by supersymmetry breaking terms. Various aspects of this model have been studied previously \\cite{herd,D37strings,previous}. The large number of light moduli fields in string inflation models \\footnote{These moduli in general could be the complex or K\\\"ahler structure moduli of the internal space. They could also be the moduli of the branes in our theory. Once the K\\\"ahler and complex structure moduli are fixed by fluxes \\cite{GVW,DRS,GKP}, they could be the remaining brane moduli. In this paper, since we are not fixing any of the moduli, the light fields could come from all the above moduli.}, however, leads to a new danger: light fields which are not the inflaton are potential entropy modes and can give rise to entropy fluctuations. The ``curvaton'' \\cite{Mollerach,SY,LM,Moroi,LW,Sloth} and ``modulated reheating'' \\cite{DGZ,Lev,MR,Uzan,Vernizzi} scenarios are examples of this effect. Non-parametric generation of fluctuations from a global symmetry breaking (also involving an key isocurvature field) has been investigated in \\cite{KRV,Matsuda}. As pointed out in \\cite{BaVi,FB2}, these entropy modes may undergo parametric instability during the initial stages of reheating. In particular, this instability occurs on super-Hubble (but sub-horizon) scales. This is possible \\cite{FB1}since the background fields carry the causal information to scales larger than the Hubble radius. The entropy fluctuations, in turn, will seed a curvature mode (which we will call ``secondary'' mode). There are two field scalar field toy models in which this secondary curvature mode dominates over the primary mode, the pure adiabatic linear perturbation theory mode. In fact \\cite{Zibin} the secondary mode can grow to be larger than unity before back-reaction shuts off the parametric instability which drives the growth of the entropy fluctuation. In this case, the model is ruled out by the observed (small) amplitude of curvature fluctuations on large cosmological scales. The D3/D7 brane inflation model is a prototypical example in which there is an entropy mode which can undergo resonance during the initial stages of reheating, the preheating \\cite{JB1,KLS1,JB2,KLS2} phase. In this paper, we study the growth of this entropy mode and calculate the magnitude of the induced secondary curvature fluctuations. It turns out that for reasonable particle physics parameters, the secondary fluctuations remain in the linear regime at the end of the reheating phase. However, their amplitude can be (in the absence of back-reaction effects) larger than the amplitude of the primary mode. This would necessitate changing the parameters of the model in order to reproduce the observed magnitude of the large-scale cosmological perturbations. Note that a similar conclusion was recently found \\cite{Larissa} in another model of brane inflation, the KKLMMT model \\cite{KKLMMT}. In that model, it had already been observed that due to their large phase space enhancement, second order fluctuations may give a dominant contribution \\cite{BC2}. We should emphasize that the effect we are calculating in this paper is an effect which arises in linear cosmological perturbation theory. Both the linear effects described here and the quadratic processes discussed in \\cite{BC2} and elsewhere (e.g. \\cite{Losic,Patrick}) can be operational at the same time. In fact, the enhanced growth of the linear perturbations which is the focus of this work will increase the strength of the quadratic processes. We expect that effects similar to those discussed in \\cite{BC2} will also arise in the D3/D7 system. The structure of this paper is as follows: we first revisit the calculation of the amplitude of the spectrum of cosmological perturbations in the D3/D7 inflationary model based on the magnitude of the primary adiabatic mode alone. Next, we give an approximate but analytical study of reheating in the D3/D7 brane inflation scenario. The initial stages of reheating are governed by a tachyonic decay channel. The same tachyonic resonance channel leads to exponential growth of super-Hubble entropy modes, which then in turn induce a secondary curvature mode. The last major section of this paper focuses on the determination of the amplitude of this mode. Our analysis of the reheating phase provides results which are interesting in their own right. A new aspects of our analysis is that we allow for an initial offset in the value of the ``waterfall\" field. Such an offset could arise as a consequence of the back-reaction of long wavelength \\footnote{Long means longer than the Hubble radius.} fluctuations of the matter fields. Such an offset will suppress the formation of domains on small scales. If the offset were sufficiently large, it could prevent the onset of the tachyonic instability. However, if we take the offset to be generated by the abovementioned fluctuations, we find that for most of the realistic space of parameters of the D3/D7 inflationary model, the resulting values of the field lie in the configuration space region for which the tachyonic resonance channel \\cite{tachyonic} is effective. Thus, the initial stages of the transfer of the energy in the inflaton field to matter fluctuations will proceed via the tachyonic instability in a time interval which is less than the characteristic time it takes for the matter fields to oscillate about their minima. A substantial part of the initial inflaton energy is converted to localized nonlinear fluctuations during the period of tachyonic resonance. Neglecting the back-reaction of these fluctuations on the homogeneous background, we see that a fraction of order unity of the initial inflaton energy density still remains in the background when the tachyonic instability shuts off. We are interested in demonstrating analytically that the transfer of energy from the inflaton to fluctuations is effective in rapidly draining almost all of the energy from the homogeneous fields. The dominant energy conversion processes are those related to the interactions of the nonlinear fluctuations. These have been studied numerically in great detail in \\cite{tachyonic}. Since our aim is to demonstrate analytically that within less than a Hubble time the energy can be effectively drained from the homogeneous background, we will neglect the nonlinear interactions and focus on the evolution of the homogeneous component of the fields once the tachyonic resonance shuts off (due to our approximation scheme, we will thus only obtain an upper bound for the decay time). We show that the decay of the residual homogeneous component of the inflaton field occurs via a parametric resonance instability \\cite{JB1,KLS1,JB2,KLS2}. In spite of the fact that the resonance turns out to be of narrow-band type, it leads to an efficient conversion of most of the energy from the background inflaton field to matter fluctuations. One of the advantages of the D3/D7 brane inflation model is that the matter fields are directly associated with the branes whose separation provides the inflaton field\\footnote{Although not directly related to the main theme of the paper, we would like to point out that some recent papers \\cite{beasly} have shown how to get standard model with chiral fermions in a D3/D7 set-up from F-theory. Of course one need to change our background manifold to a certain Del-Pezzo surface to accomodate the standard model. It would be interesting to realise D3/D7 inflationary dynamics with this manifold.}. As a consequence, there is a direct coupling between the inflaton field and the matter fields, rendering it easy, as we will see, to obtain efficient reheating. This is a feature the D3/D7 brane inflation model shares with the D4/D6 model in which reheating was analyzed in \\cite{Easson} and with the brane-antibrane inflation models in which reheating was studied in \\cite{CFM,BC} (see also \\cite{TM}), but contrasts with the situation in inflationary models in which the inflaton dynamics and the standard model fields live in different throats, in which case efficient reheating is more difficult to obtain (see e.g. \\cite{BBC,KY,DSU,FMM,CT,PL}). The outline of this paper is as follows. In the following section, we review the D3/D7 brane inflationary model, with particular emphasis on the form of the potential for the inflaton field. In Section 3 we discuss the vacuum manifold of the model, and in Section 4 we revisit the calculation of the spectrum of cosmological perturbations. Section 5 focuses on the dynamics of the inflaton and of the other relevant scalar fields of the model before the onset of reheating, introduces the conditions for effectiveness of tachyonic resonance and demonstrates that for most of the relevant parameter space of the theory, the tachyonic resonance condition is satisfied. However, the tachyonic instability turns off before most of the energy of the inflaton field is released as matter fluctuations. To obtain an upper bound on the time it takes for the inflaton energy stored in the residual oscillations to further dissipate, we study the later dynamics of the homogeneous field components which proceeds via the less efficient parametric resonance channel (Section 6). Section 7 is devoted to the computation of the secondary curvature fluctuation mode, the mode induced by the entropy fluctuations. ", "conclusions": "We have studied reheating and structure formation in the D3/D7 brane inflation model of \\cite{DHHK}, with particular emphasis on the tachyonic instability of super-Hubble scale entropy fluctuations. These entropy perturbations induce a curvature fluctuations which we call ``secondary''. We have found that under certain conditions these secondary fluctuations are larger than the primordial adiabatic ones which have been considered in the past. They do, however, remain in the regime of applicability of linear perturbation theory. This would imply that the parameters of the model have to be changed compared to what is usually assumed in order to obtain agreement with the observed amplitude of the large-scale curvature fluctuations. Along the way, we have given an extensive analytical analysis of the reheating mechanism in the D3/D7 brane inflation model\\footnote{Our analytical analysis is complementary to the numerical work of \\cite{tachyonic}. We need to make certain approximations and neglect some effects, as discussed in the text. On the other hand, numerical work can only handle a limited range of scales. We are interested in large cosmological scales, whereas the basic physical scale of the system is microphysical. It is not clear that numerical work which is sensitive to the microphysical scale can make reliable predictions for questions involving cosmological scales.}. In the low energy field theory limit, the dynamics of this system is a special case of supersymmetric hybrid inflation. We have taken into account the quantum fluctuations in the ``waterfall'' field $\\phi_+$. These play an important role for both the reheating dynamics and for the generation of entropy fluctuations. For COBE-normalized values of the string theory parameters and for reasonable values of the string coupling constant we have found that, including the effects of the above mentioned quantum fluctuations, the initial stages of reheating occur via a tachyonic instability. This instability, however, shuts off quite early, leaving some of the initial inflaton energy in the background fields. To obtain an upper bound on the time it takes to drain most of the residual energy from the background homogeneous fields, we have neglected the interactions of the nonlinear fluctuations produced during the initial phase of tachyonic decay, and focused on the residual homogeneous field dynamics. This later dynamics proceeds by the parametric resonance instability of \\cite{JB1,KLS1,JB2,KLS2}. In fact, except for at the onset of the reheating process, the system is in the narrow-band region of parameters. Nevertheless, the reheating is sufficiently efficient to convert a substantial fraction of the inflaton energy into matter fluctuations. We note that efficient reheating in the D3/D7 brane inflation model is easier to achieve than in multi-throat inflation models since the matter fields are directly coupled to the inflaton in the form of strings stretching between the two branes whose separation constitutes the inflaton. In our work, we have neglected the issue of moduli stabilization. This is a very important caveat to our analysis. Moduli stabilization in our model has been considered in the first reference of \\cite{previous} and more recently in \\cite{BBDD}. We plan to study the implications of the corrections to the potential induced by moduli stabilization on the dynamics of reheating in a followup paper. We have also simplified the dynamics of the background fields. Namely, we have considered in-phase oscillations of the two background fields $S$ and $\\phi_+$, and we have neglected their phases. These phases provide extra low mass entropy modes. It would be interesting to consider their excitation during reheating." }, "0801/0801.4438_arXiv.txt": { "abstract": "The proposed design for PILOT is a general-purpose, wide-field ($1\\dg$) 2.4m, f/10 Ritchey-Chr\\'etien telescope, with fast tip-tilt guiding, for use $0.5-25\\mum$. The design allows both wide-field and diffraction-limited use at these wavelengths. The expected overall image quality, including median seeing, is $0.28-0.3''$ FWHM from $0.8-2.4\\mum$. Point source sensitivities are estimated. ", "introduction": "The free atmospheric conditions at Dome C are known to be exceptional. The seeing (above $\\sim$ 30m), coherence time, isoplanatic angle, infrared sky emission, water vapour absorption and telescope thermal emission are all better than at the best mid-latitude sites (Lawrence {\\em et al.\\/} \\cite{Lawrence}, Agabi {\\em et al.\\/} \\cite{Agabi}, Walden {\\em et al.\\/} \\cite{Walden}). PILOT (the Pathfinder for an International Large Optical Telescope) is intended to show that we can fully utilise these conditions for optical/infra-red astronomy. It is also intended to demonstrate that large optical telescopes can be built and operated in Antarctica; to fully characterise the possibilities for adaptive optics; and to perform cutting-edge science in its own right. A design study by the AAO and UNSW is currently underway, with completion mid-2008. As part of that study, we are actively seeking partners in the PILOT project, and input into the design and the scientific requirements \\begin{figure} \\begin{center} \\includegraphics[width=9.5cm,angle=-90]{saunders1_fig1.ps} \\end{center} \\caption{Optical layout for PILOT, for wide-field optical ($r$-band) use with $1\\dg$ field (above) and wide-field infra-red ($K_{dark}$-band) use with $15'$ field and cold stop (below). The spot diagrams (on right) include the Airy disc and a scale bar of 18$\\mum$/$0.15''$ (above) and 72$\\mum$/$0.62''$ (below). The optical design is diffraction limited above 500nm.} \\end{figure} ", "conclusions": "" }, "0801/0801.1501_arXiv.txt": { "abstract": "Using {\\it Spitzer} archival data from the SAGE (Surveying the Agents of a Galaxy's Evolution) program, we derive the Cepheid period-luminosity (P-L) relation at $3.6$, $4.5$, $5.8$ and $8.0$ microns for Large Magellanic Cloud (LMC) Cepheids. These P-L relations can be used, for example, in future extragalactic distance scale studies carried out with the {\\it James Webb Space Telescope}. We also derive Cepheid period-color (P-C) relations in these bands and find that the slopes of the P-C relations are relatively flat. We test the nonlinearity of these P-L relations with the $F$ statistical test, and find that the $3.6\\mu \\mathrm{m}$, $4.5\\mu \\mathrm{m}$ and $5.8\\mu \\mathrm{m}$ P-L relations are consistent with linearity. However the $8.0\\mu \\mathrm{m}$ P-L relation presents possible but inconclusive evidence of nonlinearity. ", "introduction": "The Cepheid period-luminosity (P-L) relation is an important component in extragalactic distance scale and cosmological studies. The most widely used P-L relations in the literature are obtained from Large Magellanic Cloud (LMC) Cepheids in optical $BVI$ \\citep[e.g., see][]{mad91,tan99,uda99a,san04,kan06} and near infra-red (NIR) $JHK$ \\citep[e.g., see][]{mad91,gie98,gro00,nik04,per04,nge05} bands. In addition, there are also some LMC P-L relations obtained for MACHO $V_{MACHO}R_{MACHO}$ \\citep{nik04,nge05} and EROS $V_{EROS}R_{EROS}$ \\citep{bau99} bands. Recently, \\citet{nge07} applied a semiempirical approach to derive the LMC P-L relation in Sloan $ugriz$ bands: still within optical and/or NIR regimes. Therefore, Cepheid P-L relations are well-developed for wavelengths ranging from optical to NIR. In contrast, there are currently no P-L relations available for wavelengths longer than the $K$ band in the literature. The main motivation for having a P-L relation at these longer wavelengths is in order to apply it in future extragalactic distance scale studies. The {\\it Near Infrared Camera (NIRCam)} and the {\\it Mid-Infrared Instrument (MIRI)}, that are scheduled to be installed on the {\\it James Webb Space Telescope (JWST)}, will operate in the NIR and mid-infrared: a wavelength range of 0.6-5 microns and 5-27 microns, respectively\\footnote{See the links given in {\\tt http://www.stsci.edu/jwst/instruments/}}. It is possible that extragalactic Cepheids will be discovered and/or (re-)observed (for those galaxies that were observed by the $HST\\ H_0$ Key Project) by the {\\it JWST}. Hence, P-L relations at longer wavelengths are needed in order to use {\\it JWST} to derive a Cepheid distance to these galaxies. In this Paper, we derive the LMC P-L relations at $3.6$, $4.5$, $5.8$ and $8.0$ microns from the {\\it Spitzer} archival data. As far as we are aware, this is the first time such a relation has been derived. In addition to the P-L relations, we can also derive period-color (P-C) relations and construct color-color plots and the color-magnitude diagrams (CMD) for LMC Cepheids in these {\\it Spitzer} bands. In Section 2 we discuss our data selection. In Section 3 we present our analysis and results for the P-L relations. Section 4 we show the P-C relations, the CMD and the color-color plot for the Cepheids in our sample. Our conclusion is given in Section 5. Extinction is ignored in this Paper because it is expected to be negligible in the {\\it Spitzer's IRAC} bands (hereafter {\\it IRAC} band). ", "conclusions": "In this Paper, we derive P-L relations for LMC Cepheids in {\\it IRAC} $3.6$, $4.5$, $5.8$ and $8.0$ microns bands. These P-L relations can be potentially applied to future extragalactic distance scale studies with, for example, the {\\it JWST}. The data are taken from the {\\it Spitzer's} archival database from the SAGE program. After properly removing the outliers, the fitted P-L relations are presented in Table \\ref{tab1}. We have tested the P-L relations with various period cuts and found that our results are insensitive to period cuts up to $\\log P_{cut} \\sim 0.65$. We also argue that the random phase corrections may not be important for {\\it IRAC} band P-L relations. When comparing P-L relations from $B$ to $8.0\\mu \\mathrm{m}$ bands, the slope of the P-L relation appears to be the steepest around $K$ band to $3.6\\mu \\mathrm{m}$ band, while the dispersion of the P-L relation reaches a minimum between those bands. The shallower slopes and larger P-L dispersions in the $5.8\\mu \\mathrm{m}$ and $8.0\\mu \\mathrm{m}$ band are in contrast to the theoretical expectation. This could be due to the smaller number of Cepheids and larger measurement errors toward the faint end in these two bands. We also test the nonlinearity of the P-L relations in the {\\it IRAC} band using the $F$ statistical test. As expected, the $F$-test results show that the P-L relations are linear in $3.6$, $4.5$ and $5.8$ microns band, but the $8.0\\mu \\mathrm{m}$ P-L relation is found to be nonlinear. However, the nature of the nonlinear $8.0\\mu \\mathrm{m}$ P-L relation is still inconclusive. For the P-C relations, it was found that the slopes of the P-C relation are relatively flat in the {\\it IRAC} bands. Finally, the LMC Cepheids show a well-defined instability strip in the CMD and clustered in a small region in the color-color plot. This can potentially be used to identify Cepheids observed in the {\\it IRAC} bands. Even though there may be some associated problems for the $8.0\\mu \\mathrm{m}$ P-L relation, the $3.6\\mu \\mathrm{m}$, $4.5\\mu \\mathrm{m}$ and perhaps the $5.8\\mu \\mathrm{m}$ P-L relations can still be used in future distance scale studies." }, "0801/0801.3732_arXiv.txt": { "abstract": "{} {We report on the production of a large area, shallow, sky survey, from \\xmm slews. The great collecting area of the mirrors coupled with the high quantum efficiency of the EPIC detectors have made \\xmm the most sensitive X-ray observatory flown to date. We use data taken with the EPIC-pn camera during slewing manoeuvres to perform an X-ray survey of the sky. } { Data from 218 slews have been subdivided into small images and source searched. This has been done in three distinct energy bands; a soft (0.2-2 keV) band, a hard (2-12 keV) band and a total \\xmm band (0.2-12 keV). Detected sources, have been quality controlled to remove artifacts and a catalogue has been drawn from the remaining sources.} {A 'full' catalogue, containing 4710 detections and a 'clean' catalogue containing 2692 sources have been produced, from 14\\% of the sky. In the hard X-ray band (2-12 keV) 257 sources are detected in the clean catalogue to a flux limit of $4\\times10^{-12}$ ergs s$^{-1}$ cm$^{-2}$. The flux limit for the soft (0.2-2 keV) band is $6\\times10^{-13}$ ergs s$^{-1}$ cm$^{-2}$ and for the total (0.2-12 keV) band is $1.2\\times10^{-12}$ ergs s$^{-1}$ cm$^{-2}$. The source positions are shown to have an uncertainty of 8\\arcsec (1~$\\sigma$ confidence).} {} ", "introduction": "There is a strong tradition in X-ray astronomy of using data taken during slewing manoeuvres to perform shallow surveys of the sky. The Einstein (\\cite{elvis}), Exosat (\\cite{reynolds}) and RXTE (\\cite{revnivtsev}) slew surveys all provide a useful complement to dedicated all-sky surveys such as ROSAT (\\cite{voges}) and HEAO-1 (\\cite{piccinotti}) and the smaller area, medium sensitivity ASCA survey (\\cite{ueda}) and pencil-beam \\xmm (\\cite{hasinger}) and Chandra (\\cite{brandt}) deep looks. It was long recognised that XMM-Newton (\\cite{jansen}) with its great collecting area, efficient CCDs, wide energy band and tight point-spread function (PSF) has the potential to make an important contribution to our knowledge of the local universe from its slew data. Early estimates (\\cite{Lumb98}; \\cite{jonesandlumb}), based on an expected slewing speed of 30 degrees per hour, and a slightly lower background level than that actually encountered in orbit, predicted a 0.5--2 keV flux limit of $2\\times10^{-13}$ erg cm$^{-2}$s$^{-1}$. While the chosen in-orbit slew speed of 90 degrees per hour reduces the sensitivity, initial assessments of the data showed that the quality of the data is good and that many sources are detected (\\cite{freyberg}). A review of properties shows that the \\xmm slew survey compares favourably with other large area surveys in terms of depth and positional accuracy (Table 1). During slews all the three imaging EPIC cameras take data in the observing mode set in the previous pointed observation and with the Medium filter in place. The slew speed of 90 degrees per hour combined with the slow readout time of the MOS detectors (2.6s; \\cite{turner}) means that sources appear as long streaks in the MOS cameras but are well formed in the fast observing modes of the pn camera (\\cite{struder}). For this reason, only the EPIC-pn data have been analysed. In this paper we present a catalogue drawn from slews taken between revolutions 314 and 978 covering a sky region of 6240 square degrees. The main properties of the slew survey discussed in this paper are given in Table 2. \\begin{table} \\caption{Properties of large area X-ray surveys} \\label{table:SurveySumm} % \\begin{center} \\begin{tabular}{l c c c c} % \\hline\\hline % Satellite & Energy range & Coverage$^{a}$ & Flux lim. & Position \\\\ % & (keV) & \\% of sky & $^{b}$ & error \\\\ % \\hline % RASS & 0.2-2.4 & 92 & 0.03 & 12\\arcsec \\\\ % Einstein slew & 0.2-3.5 & 50 & 0.3 & 1.2\\arcmin \\\\ {\\bf XMM slew (soft)} & {\\bf 0.2-2} & {\\bf 14} & {\\bf 0.06} & {\\bf 8\\arcsec} \\\\ \\hline % EXOSAT slew & 1-8 & 98 & 3 & 20\\arcmin \\\\ HEAO-1/A2 & 2-10 & 100 & 3 & 60\\arcmin \\\\ % RXTE slew & 3-20 & 95 & 1.8 & 60\\arcmin \\\\ {\\bf XMM slew (hard)} & {\\bf 2-12} & {\\bf 14} & {\\bf 0.4} & {\\bf 8\\arcsec} \\\\ \\hline % \\end{tabular} \\\\ \\end{center} $^{a}$ The \\xmm slew sky coverage has been computed by adding the area contained in all of the images used in source searching with an exposure time greater than 1 second. \\\\ $^{b}$ Flux limit, units of $10^{-11}$ ergs $s^{-1}$ cm$^{-2}$ \\\\ \\end{table} \\begin{table} \\caption{Properties of the \\xmm slew survey} \\label{table:slewprops} % \\begin{center} \\begin{tabular}{l c c c} % \\hline\\hline % Property & \\multicolumn{3}{c}{Range (keV)}\\\\ % & 0.2--2 & 2--12 & 0.2--12 \\\\ % \\hline % Observing time (s) & $5.4\\times10^{5}$ & $5.4\\times10^{5}$ & $5.4\\times10^{5}$ \\\\ Mean exposure time$^{a}$ & 6.2 s & 6.0 s & 6.1 s\\\\ Total number of photons & $1.6\\times10^{6}$ & $2.2\\times10^{6}$ & $3.8\\times10^{6}$ \\\\ Mean background$^{b}$ & 0.09 & 0.14 & 0.23\\\\ Median source count rate$^{c}$ & 0.68 & 0.90 & 0.81\\\\ Limiting source flux$^{d}$ & $6.0\\times10^{-13}$ & $4.0\\times10^{-12}$ & $1.2\\times10^{-12}$\\\\ Num. sources (full cat)$^{e}$ & 2606 & 692 & 3863\\\\ Num. sources (clean cat)$^{e}$ & 1874 & 257 & 2364\\\\ \\hline % \\end{tabular} \\\\ \\end{center} $^{a}$ The mean exposure time, after correcting for the energy-dependent vignetting.\\\\ $^{b}$ cnts/arcmin$^{2}$. \\\\ $^{c}$ cnts/second. \\\\ $^{d}$ ergs/s/cm$^{2}$. Based on a detection of 4 photons from a source passing near the detector centre (exp. time = 10s), with a power-law spectrum of slope 1.7 and galactic absorption of 3$\\times10^{20}$ atoms cm$^{-2}$. \\\\ $^{e}$ The number of sources flagged as good in the full and clean catalogues (see text). \\\\ \\end{table} Slews have been source searched down to a likelihood threshold of 8, which after manual rejection of false detections gives 4710 candidate sources. Using simulations we have been able to identify a subset of high significance sources, with a likelihood threshold dependent upon the background conditions, that gives 2692 sources with a spurious fraction of $\\sim4\\%$ (the \"clean\" catalogue). Of these, 2621 are from unique sources. In the hard (2--12 keV) band the clean catalogue contains 257 sources (253 unique) of which $\\sim9\\%$ are expected to be due to statistical fluctuations. The slew catalogue and accompanying images and exposure maps have been made available through the XMM Science Archive (XSA) as a queryable database and as FITS files\\footnote{The catalogue was initially released on May 3 2006. In this paper we discuss the updated version released in October 2006.}. A summary of scientific highlights from the slew survey has been published in \\cite{Read06}. ", "conclusions": "The \\xmm slew data represent the deepest near all-sky hard band X-ray survey made to date, while the soft band survey is comparable with the RASS. The source density, from the clean catalogue, is $\\sim0.45$ per square degree and $\\sim70\\%$ have plausible identifications. The \\xmm slew survey catalogue will continue to grow as the mission continues and it is expected that a sky coverage in excess of 50\\% will eventually be achieved. With the excellent attitude reconstruction this will leave a powerful legacy for future variability studies in the soft and hard X-ray bands." }, "0801/0801.0605.txt": { "abstract": "% context heading (optional) % {} leave it empty if necessary {Over the last decade, several groups of young (mainly low-mass) stars have been discovered in the solar neighbourhood (closer than $\\sim$100 pc), thanks to cross-correlation between X-ray, optical spectroscopy and kinematic data. These young local associations -- including an important fraction whose members are Hipparcos stars -- offer insights into the star formation process in low-density environments, shed light on the substellar domain, and could have played an important role in the recent history of the local interstellar medium.} % aims heading (mandatory) {To study the kinematic evolution of young local associations and their relation to other young stellar groups and structures in the local interstellar medium, thus casting new light on recent star formation processes in the solar neighbourhood.} % methods heading (mandatory) {We compiled the data published in the literature for young local associations. Using a realistic Galactic potential we integrated the orbits for these associations and the Sco-Cen complex back in time.} % results heading (mandatory) {Combining these data with the spatial structure of the Local Bubble and the spiral structure of the Galaxy, we propose a recent history of star formation in the solar neighbourhood. We suggest that both the Sco-Cen complex and young local associations originated as a result of the impact of the inner spiral arm shock wave against a giant molecular cloud. The core of the giant molecular cloud formed the Sco-Cen complex, and some small cloudlets in a halo around the giant molecular cloud formed young local associations several million years later. We also propose a supernova in young local associations a few million years ago as the most likely candidate to have reheated the Local Bubble to its present temperature.} % conclusions heading (optional), leave it empty if necessary {} ", "introduction": "The first moving groups of stars were discovered early on in the history of Galactic dynamics. As early as 1869, R.A. Proctor published a work in which he identified a group of stars around the Hyades open cluster that was moving together through the Galaxy. He also found another 5 comoving stars in the Ursa Major constellation. In the 1960s, O.J. Eggen suggested the existence of a {\\it Local Association} (see for example Eggen \\cite{Eggen61}, \\cite{Eggen65a}, \\cite{Eggen65b}) formed by a group of young stars with approximately the same spatial velocity (also referred as the {\\it Pleiades moving group}). Eggen's Local Association included the nearest bright B-type stars, stars in the Pleiades, the $\\alpha$ Perseus and IC2602 clusters, and the stars belonging to the Sco-Cen complex. However, it was difficult to defend a unitary view of the group due to the wide range of ages ($\\la$100 Myr) and widespread spatial distribution ($\\sim$300 pc) of its constituent stars. At the same time as Eggen was studying his Local Association, the first X-ray detectors were installed in rockets and launched into space. This led to the discovery of the diffuse soft X-ray background (SXRB). The anticorrelation observed between the SXRB and the HI column density was rapidly interpreted as evidence of a local cavity in the interstellar medium (ISM) filled by an X-ray emitting plasma, which became known as the {\\it Local Bubble} (LB). The launch of the ROSAT satellite in 1990 allowed the LB to be studied in more detail. The opening of the X-ray window in the sky and observations of stars belonging to clusters with known ages also led to the realisation that X-ray emission persists in young stars for a period of the order of 100 Myr. An important fraction of these young, X-ray emitting stars are T Tauri stars. These very young ($\\la$10 Myr) stars also exhibit excess IR emission (due to the presence of nearby heated dust particles forming disks or envelopes) as well as UV line and continuum emission produced by the accretion of the surrounding gas and dust. When the stars reach the age of $\\sim$10 Myr, both IR emissions and optical activity decline considerably. X-ray activity then becomes the basis for ascertaining youthfulness. The ROSAT All-Sky Survey (RASS) detected more than 100,000 X-ray sources at keV energies (Voges et al. \\cite{Voges00}). Using a kinematic approach and searching for groups of comoving stars allows us to determine which of them are T Tauri stars. This greatly reduces the number of candidate stars that must be observed spectroscopically to confirm their youthfulness. In this way, several young stellar associations have been discovered within 100 pc of Earth during the last decade (see Jayawardhana \\cite{Jayawardhana00}, and Zuckerman \\& Song \\cite{Zuckerman04b} for a recent review). The stars belonging to these young local associations (YLA) have ages ranging from a few million to several tens of millions of years. Due to their proximity to us, these stars are spread over a large area of the sky (up to several hundred square degrees) making it difficult to identify the associations as single entities. Furthermore, they are far away from molecular clouds or star forming regions (SFR). This is why they remained unnoticed until recently. These young stars offer insights into the star formation process in low-density environments, which is different from the dominant mechanism observed in denser SFR. The discovery of these YLA has also contributed to substellar astrophysics, since dozens of dwarfs have been identified in them. For instance, it has been confirmed that isolated brown dwarfs can form in these low-density environments (see for example Webb et al. \\cite{Webb99}; Lowrance et al. \\cite{Lowrance99}; Chauvin et al. \\cite{Chauvin03}). YLA also provide important clues regarding recent star formation in the solar neighbourhood (since they contain the youngest stars near the Sun) and its effect on the local ISM. In Sect. \\ref{sect.LB+YLA} of this paper we review current understanding of the LB and the Sco-Cen complex and present published data for the nearly 300 stars belonging to YLA. In Sect. \\ref{sect.orbits} we present our method for orbit integration, which makes use of a Galactic potential that includes the general axisymmetric potential and the perturbations due to the Galaxy's spiral structure and the central bar. This leads us to study, in Sect. \\ref{sect.origin}, the origin and evolution of the local structures and, in Sect. \\ref{sect.sce}, to propose a scenario for recent star formation in the solar neighbourhood. This scenario explains the origin of the Sco-Cen complex and YLA through the collision of a parent giant molecular cloud (GMC) with the Sagittarius-Carina spiral arm. Finally, the conclusions of our work are summarised in Sect. \\ref{sect.conc}. % %__________________________________________________________________ ", "conclusions": "\\label{sect.conc} This paper studies the kinematic evolution of the Sco-Cen complex and the so-called young local associations (YLA). It makes use of most of the astrophysical data published in the literature for all the known members of YLA (more than 200 stellar systems). This information appears in Appendix \\ref{sec.app} and can also be accessed via a webpage\\footnote{http://www.am.ub.es/$\\sim$dfernand/YLA/}. The study of the orbits integrated back in time for all these associations allows us to propose a scenario for recent star formation in the solar neighbourhood (Sect. \\ref{sect.sce}) and a possible link between these associations and the origin and/or evolution of the LB (Sect. \\ref{sect.YLA+LB}). In our scenario, the oldest Sco-Cen associations were brought about by the impact of the inner spiral arm against a giant molecular cloud. YLA were born later (and outside Sco-Cen) due to the shock fronts of the most massive supernovae belonging to the LCC or UCL associations. Our results suggest that a YLA is the most likely place to have harboured the supernova that reheated the LB a few million years ago. As seen in this paper, the recent discovery of a set of YLA, together with the use of appropriate tools (mainly, the integration back in time of orbits), sheds light on several apparently unrelated topics in astrophysics. These topics range from star formation mechanisms for low-mass stars distant from star forming regions, to the recent history of the LB, and include the origin of the Sco-Cen stellar complex and its possible independence from the Gould Belt. The possible discovery in the coming years of new members of these associations, or even of new associations, will be very useful in confirming the results obtained in this work." }, "0801/0801.2391_arXiv.txt": { "abstract": "We observed the first known very high energy (VHE) $\\gamma$-ray emitting unidentified source, TeV J2032+4130, for 94 hours with the MAGIC telescope. The source was detected with a significance of 5.6$\\sigma$. The flux, position, and angular extension are compatible with the previous ones measured by the HEGRA telescope system five years ago. The integral flux amounts to (4.5$\\pm$0.3$_{\\rm stat}\\pm$0.35$_{\\rm sys}$)$\\times$10$^{-13}$ ph~cm$^{-2}$~s$^{-1}$ above 1~TeV. The source energy spectrum, obtained with the lowest energy threshold to date, is compatible with a single power law with a hard photon index of $\\Gamma$=-2.0$\\pm$0.3$_{\\rm stat}\\pm$0.2$_{\\rm sys}$. ", "introduction": "The TeV source J2032+4130 (Aharonian et al. 2002) was the first unidentified very high energy (VHE) $\\gamma$-ray source, and also the first discovered extended TeV source, likely to be Galactic. Intensive observational campaigns at different wavelengths have been carried out on TeV J2032+4130. Butt et al. (2003) presented an analysis of the CO, HI, and infrared emissions, together with first observations by {\\it Chandra} (5 ksec) and a reanalysis of VLA data. These observations showed that the TeV source region is positionally coincident with an outlying group of stars (from the Cygnus OB2 core), although they failed to identify a counterpart. Mukherjee et al. (2003) analyzed the same {\\it Chandra} data and provided optical follow-up observations of several of the brightest X-ray sources, confirming that most were either O stars or foreground late-type stars. A deeper {\\it Chandra} observation (50 ksec, Butt et al. 2006), found hundreds of star-like sources and yet no diffuse X-ray counterpart emission. A deep ($\\sim$ 50~ksec) \\textit{XMM-Newton} exposure has also been obtained (Horns et al. 2007). After the subtraction of the contribution of known sources from the data, an extended X-ray emission region with a FWHM size of $\\sim$~12 arcmin was reported. The centroid of the emission is co-located with the position of TeV J2032+4130 and was proposed as the counterpart of the TeV source. The question whether the result reported by Horns et al. can be interpreted as a truly diffuse background, or it could be a result of unresolved X-ray sources, remains disputable. Paredes et al. (2007) and Mart\\'i et al. (2007) have provided deep radio observations covering the TeV J2032+4130 vicinity using the Giant Metrewave Radio Telescope and discovered a population of radio sources, some in coincidence with X-ray detections by Butt et al. (2006) and with optical/IR counterparts. At least three of these sources are non-thermal, and one has a hard X-ray energy spectrum. They found extended non-thermal diffuse emission in the radio band apparently connecting with one or two radio sources. It is yet to be determined if one or more of these sources is similar to some of the known $\\gamma$-ray binaries (e.g., Aharonian et al. 2006, Albert et al. 2006a). Several theoretical explanations for the TeV emission from J2032+4130 have been given. Among them, those related with extragalactic counterparts, e.g., a radiogalaxy (Butt et al. 2006) or a proton blazar (Mukherjee et al. 2003), face the difficulty of explaining the extended appearance of the source. Gamma-ray production in hypothetical jet termination lobes of Cyg X-3 was explored (Aharonian et al. 2002), but the putative northern lobe of Cyg X-3 (now considered a mere thermal HII region, Mart\\'i et al. 2006) is far from the location of the TeV source. A yet unknown pulsar wind nebula (PWN) was proposed by Bednarek (2003), although no clear PWN signal was observed. A distant microquasar was proposed by Paredes et al. (2007), perhaps related with one of the X-ray/radio sources they discovered. If such an association is accepted, the extension of the source could be explained by the diffusion of accelerated particles into a hypothetical nearby molecular enhancement (see Bosch-Ramon et al. 2005). Torres et al. (2004) and Domingo-Santamar\\'ia \\& Torres (2006) studied the relationship between the TeV emission and the known massive stars in the area, through the interaction of relativistic protons with wind ions. The distribution of stars in the neighborhood favors this interpretation (Butt et al. 2006). An explanation involving the excitation of giant dipole resonances of relativistic heavy nuclei in radiation dominated environments has also been suggested (Anchordoqui et al. 2007). ", "conclusions": "MAGIC observations confirm the location of TeV J2032+4130 found by HEGRA. The MAGIC observation shows an extended source with a significance of 5.6$\\sigma$. We find a steady flux with no significant variability within the three year span of the observations (with the flux being at a similar level of the HEGRA data of the period 2002-2005). We also present the source energy spectrum obtained with the lowest energy threshold to date, which, within errors, is compatible with a single power law." }, "0801/0801.3324.txt": { "abstract": "A numerical approach is considered for spherically symmetric spacetimes that generalize Lema\\^\\i tre--Tolman--Bondi dust solutions to nonzero pressure (``LTB spacetimes''). We introduce quasi--local (QL) variables that are covariant LTB objects satisfying evolution equations of Friedman--Lema\\^\\i tre--Robertson--Walker (FLRW) cosmologies. We prove rigorously that relative deviations of the local covariant scalars from the QL scalars are non--linear, gauge invariant and covariant perturbations on a FLRW formal ``background'' given by the QL scalars. The dynamics of LTB spacetimes is completely determined by the QL scalars and these exact perturbations. Since LTB spacetimes are compatible with a wide variety of ``equations of state'', either single fluids or mixtures, a large number of known solutions with dark matter and dark energy sources in a FLRW framework (or with linear perturbations) can be readily examined under idealized but non--trivial inhomogeneous conditions. Coordinate choices and initial conditions are derived for a numerical treatment of the perturbation equations, allowing us to study non--linear effects in a variety of phenomena, such as gravitational collapse, non--local effects, void formation, dark matter and dark energy couplings and particle creation. In particular, the embedding of inhomogeneous regions can be performed by a smooth matching with a suitable FLRW solution, thus generalizing the Newtonian ``top hat'' models that are widely used in astrophysical literature. As examples of the application of the formalism, we examine numerically the formation of a black hole in an expanding Chaplygin gas FLRW universe, as well as the evolution of density clumps and voids in an interactive mixture of cold dark matter and dark energy. ", "introduction": "Recent observations suggest that present cosmic dynamics is dominated by elusive sources called ``dark matter'' and ``dark energy'', the former clustered in galactic halos and the latter possibly associated with a large scale repulsive (yet unknown) interaction. A large body of phenomenological and theoretical models have been proposed to describe these sources (see \\cite{review} for a comprehensive review). The dominant model of dark matter is a collision--less gas of supersymmetric particles: ``cold'' dark matter~\\cite{review}, whereas dark energy has been described by a cosmological constant \\cite{Lambda}, as well as by quintessence scalar fields, Phantom fields, tachyons, branes (see \\cite{review}). A unified perspective is given by the Chaplygin gas~\\cite{review,chaplygin1,chaplygin2}, a single source that exhibits the expected dynamical behavior of dark matter and dark energy. There is also a large number of empiric and phenomenological descriptions of dark matter interacting with dark energy \\cite{intmix1,intmix2} (see \\cite{intmix_new} for a recent appraisal), as well as empiric dark energy ``equations of state'' that fit observations~\\cite{EOS1,EOS2}. While dark matter at the galactic scale is considered inhomogeneous, due to its association with structure formation, most research work on dark energy sources has been conducted in homogeneous Friedman--Lema\\^\\i tre--Robertson--Walker (FLRW) spacetimes and/or linear perturbations, since dark energy is assumed to have a significant effect only at a larger cosmic scale in which the universe appears to be homogeneous. However, the possibility of anisotropic or inhomogeneous dark energy (or combined dark matter and energy sources) is beginning to be discussed in the literature \\cite{chapinhom,chapinhom2,chapstar,intmix3,Mota_anis}. Given our ignorance on the fundamental properties of these sources, there is no theoretically binding reason to assume, {\\it a priori}, that no valuable new information would result from studying their interaction under inhomogeneous conditions, at least in scales associated with structure formation and gravitational clustering. Alternative proposals to the dark energy paradigm also consider a full relativistic treatment of cosmological inhomogeneities~\\cite{Inh_rew1,Inh_rew2}. Numerous articles describe the possibility that cosmic acceleration (or at least part of it) could result from the presence of inhomogeneites in photon trajectories in observations from high red--shift objects~\\cite{InhObs1, InhObs2}. Following a more theoretical perspective, various averaging formalisms~\\cite{ave1,ave2} have examined the occurrence of an effective acceleration from the so--called ``back--reaction'', associated with non--local effects that emerge as inhomogeneous sources are averaged (see \\cite{theo1,theo2} for further discussion and \\cite{ltbave,condsBR} for scalar averages in the spherical dust case). Also, an interesting connection between cosmic acceleration, back--reaction and non--trivial gradients of quasi--local energy has been proposed~\\cite{wiltshire}, which could clarify important theoretical issues (the connection between the material presented in this paper and back--reaction issues is discussed in \\cite{condsBR,QLvars}). Possible observational consequences of these theoretical alternatives to dark energy are certainly worth serious consideration~\\cite{Obs_BR} (see also \\cite{InhObs1,InhObs2} ). Even if dark energy prevails over these alternative proposals, there is no harm done (and perhaps valuable new information) in probing its behavior under inhomogeneous conditions. Ideally, the study of inhomogeneous sources should be conducted on spacetimes not restricted by simplifying symmetries, but this would require fully 3--dimensional codes of high complexity. As a compromise between linear perturbations and this type of ``realistic'' generality, we present in this article a class of inhomogeneous spherically symmetric models that can be well handled with relatively simple numerical methods. Although these models are idealized, they are non--trivial, exhibit non--linear behavior, and are general enough for testing a wide variety of physical assumptions on modern cosmological sources. By assuming a spherically symmetric Lema\\^\\i tre--Tolman--Bondi metric, we derive in section \\ref{genLTB} a class of spacetimes (``LTB spacetimes'') that generalize to nonzero pressure (and nonzero pressure gradients) the well know solutions for inhomogeneous dust associated with this metric \\cite{kras}. The most general source for this metric (in the comoving frame) is a fluid with anisotropic stresses, which will end up being interpreted as fluctuations of the isotropic pressure. However, regardless of the interpretation, we remark that the existence of pressure anisotropy is far from a drawback or shortcoming, not only because it can be associated to a number of well motivated physical effects~\\cite{anisP}, but because an inhomogenous source with only isotropic pressure is a far more idealized situation than one with anisotropic pressure. In section \\ref{1plus3} we show that LTB spacetimes can be fully characterized by a set of local covariant scalars, hence their dynamics becomes completely determined by solving the evolution equations for these scalars in the covariant fluid flow (or ``1+3'') representation~\\cite{ellisbruni89,BDE,LRS,1plus3}. By looking at the relation between energy density and the Misner--Sharp quasi--local mass--energy invariant~\\cite{MSQLM,kodama,szab,hayward1,hayward2}, we introduce in section \\ref{QLdefs} ``quasi--local'' (QL) scalar functions, which are LTB objects that satisfy FLRW dynamics. Expressing all local covariant scalars as perturbations of the QL scalars, we rewrite in section \\ref{eveqs} the 1+3 fluid flow equations of section \\ref{1plus3} as evolution equations for the QL scalars and these perturbations, which on the basis of known criteria that define a perturbation formalism~\\cite{ellisbruni89, BDE,LRS,1plus3,bardeen}, are shown in section \\ref{perturb} to be covariant, gauge--invariant, non--linear perturbations on a FLRW formal ``background'' given by the QL variables. These evolution equations are equivalent to ODE's in which radial dependency enters as a parameter (see Appendix C of \\cite{suss08})). As shown in section \\ref{EOS}, once we choose an ``equation of state'' (EOS) between the QL energy density and pressure (which are ``background'' variables), the QL evolution equations become determined. We point out that, by choosing such an EOS, correlation terms appear in the relations between local scalars, in a similar way as virial corrections to the ideal gas EOS appear in Newtonian self--gravitating systems~\\cite{saslaw1,saslaw2}. These terms can be associated with the long range nature of gravity~\\cite{Padma_GTD}. Since the fundamental physics of dark matter and dark energy sources is still unknown, we cannot rule out the possibility that this type of non--local effects could play an important dyamical role when these sources are studied under non--trivial and non--linear inhomogeneity. In section \\ref{mixtures} we apply the perturbation formalism to a fluid mixture, which can be interactive or with each component separately conserved (decoupled). This mixture description can be readily used to generalize to inhomogeneous conditions similar mixture models derived in a FRLW context in the many references quoted in \\cite{intmix1,intmix2,intmix_new}. We show in this section that scalar fields can only be compatible with LTB spacetimes in mixture sources, playing the role of the homogeneous dark energy component (coupled with inhomogeneous dark matter~\\cite{intmix3}) Coordinate choices appropriate for the numerical treatment for the evolution equations of sections \\ref{eveqs} and \\ref{mixtures} are discussed in \\ref{FLRWlike}, while in section \\ref{FLRWsubs} we show that perturbation functions are well defined as long as shell crossing singularities are absent. A possible (but not compulsory) way in which LTB spacetimes can be embedded into a FLRW background is by smoothly matching an inhomogeneous LTB section with a section of a suitable FLRW spacetime. We provide in section \\ref{tophats} the conditions for such a matching, leading to smooth and fully relativistic generalizations of the Newtonian ``top hat'' models that are widely used in the astrophysical literature~\\cite{tophats}. We examine in section \\ref{applications} two examples of the application of the formalism presented in the previous sections. In the first example, we solve numerically the evolution equations for a smooth and fully relativistic Chaplygin gas ``top hat'' model, in which a black hole is formed in an inhomogeneous comoving section of the LTB Chaplygin gas, smoothly embedded in an expanding Chaplygin gas FLRW universe. In the second example, we examine a mixture of interactive dark matter and dark energy of the type examined in \\cite{intmix1,intmix2,intmix_new}, but give the interaction a phenomenological interpretation in terms of particle creation, with dark matter particles decaying into dark energy~\\cite{intmix_new,lima}. We show how the radial profiles of dark matter density evolve from initial clumps into deep voids, thus providing a toy model for a void formation scenario in the context of this type of interactions. In the appendices we provide detailed information on how initial conditions and coordinate choices can be set up for the numerical integration of the evolution equations. ", "conclusions": "We have presented a formalism, based on quasi--local (QL) variables, describing a large class of spherically symmetric models (LTB spacetimes) as non--linear, gauge invariant and covariant perturbations of a FLRW formal background. The integration of the resulting evolution equations require numerical work, but this task can be handled with relatively simple numerical techniques. A summary of how this formalism was motivated, introduced, developed and applied is furnished in the Introduction (section II). Since LTB spacetimes are compatible with a wide variety of ``equations of state'' (EOS) and physical assumptions, these models and the formalism developed for them are useful theoretical tools to examine a large amount of cosmological sources of interest that have been studied only in a FLRW context (or linear perturbations)~\\cite{review,Lambda,chaplygin1,chaplygin2,intmix1,intmix2,intmix_new,EOS1,EOS2}. While the inhomogeneity of LTB spacetimes is certainly idealized, the resulting models are non--trivial and still exhibit non--linear behavior, and can be useful to examine important phenomena (gravitational collapse, void formation) that cannot be studied with FLRW models or their linear perturbations. As we argued in section \\ref{EOS}, the EOS that determines the evolution equations is given between the QL energy density and pressure, so that local density and pressure and pressure are related by means of correlation terms that could contain non--local information that could be important for self--gravitating sources. Since the fundamental nature of dark matter and dark energy is not known, we cannot rule the possibility of non--local effects playing an important role when we examine these sources under inhomogeneous conditions, at least in a scale smaller than 300 Mpc. We believe that LTB models can be useful to examine these and other non--linear effects, not only for dark matter or dark energy, but also in theoretical proposals that seek to explain cosmic acceleration without dark energy~\\cite{Inh_rew1,Inh_rew2,InhObs1,InhObs2,ave1,ave2,ave3,theo1,theo2,ltbave,condsBR,wiltshire}. As shown elsewhere~\\cite{condsBR,QLvars}, the formalism presented here can be applied to understand the issue of ``back--reaction'' that appears in various averaging formalisms~\\cite{ave1,ave2,ave3} aiming to explore this type of alternative explanations. An important result of this article is the smooth and fully relativistic generalization of the Newtonian ``top hat'' models (section \\ref{tophats}), which are useful toy models for structure formation and are widely used in Astrophysical literature. One of the numerical examples of application of the formalism that we provided (section \\ref{chaplygin}) is a Chaplygin gas ``top hat'' model that describes the formation of a local black hole smoothly embedded in an expanding Chaplygin gas FLRW universe. The second numerical application example is that of an interactive mixture of cold dark matter (CDM) and dark energy. By interpreting the interaction term in terms of particle creation, we obtained a model in which initial clump (or overdensity) radial profiles of CDM evolve into void profiles as CDM decays into dark energy. While this model is not ``realistic'', it serves to illustrate how LTB spacetimes can illustrate important non--linear features of dark energy/matter sources that cannot be studied within a homogeneous context (nor with linear perturbations). As shown in \\cite{intmix2,intmix_new}, mixtures of this type can be studied by means of a dynamical systems approach. This type of approach is compatible with the evolution equations of LTB spacetimes (autonomous equations), and so a generalization can be readily made of these dynamical system studies in a FLRW context. In fact, a dynamical systems study of LTB dust solutions has been achieved using the same QL variables~\\cite{suss08}. The extension of this work for LTB spacetimes with nonzero pressure is being currently undertaken." }, "0801/0801.2450_arXiv.txt": { "abstract": "{The X-ray reflection features of irradiated accretion disks around black holes enable us to probe the effects of strong gravity. We investigate to which precision the reflection signs, i.e. the iron K-line and the Comptonized hump, can be observed with Simbol-X for nearby Seyfert galaxies. The simulations presented include accurate computations of the local reprocessed spectra and modifications due to general relativistic effects in the vicinity of the black hole. We discuss the impact of global black hole parameters and of the irradiation pattern of the disk on the resulting spectra as they will be detected by the Simbol-X mission. ", "introduction": " ", "conclusions": "" }, "0801/0801.0039_arXiv.txt": { "abstract": "Anomalous X-ray pulsars (AXPs) are thought to be magnetars, which are neutron stars with ultra strong magnetic field of $10^{14}$--$10^{15}$ G. Their energy spectra below $\\sim$10 keV are modeled well by two components consisting of a blackbody (BB) ($\\sim$0.4 keV) and rather steep power-law (POW) function (photon index $\\sim$2-4). Kuiper et al.(2004) discovered hard X-ray component above $\\sim$10 keV from some AXPs. Here, we present the Suzaku observation of the AXP 1E 1841-045 at the center of supernova remnant Kes 73. By this observation, we could analyze the spectrum from 0.4 to 50 keV at the same time. Then, we could test whether the spectral model above was valid or not in this wide energy range. We found that there were residual in the spectral fits when fit by the model of BB + POW. Fits were improved by adding another BB or POW component. But the meaning of each component became ambiguous in the phase-resolved spectroscopy. Alternatively we found that NPEX model fit well for both phase-averaged spectrum and phase-resolved spectra. In this case, the photon indices were constant during all phase, and spectral variation seemed to be very clear. This fact suggests somewhat fundamental meaning for the emission from magnetars. ", "introduction": "Anomalous X-ray pulsars (AXPs) are thought to be magnetars, which are strongly magnetized ($\\sim 10^{14}$ -- $10^{15}$ G) neutron stars with emissions powered by the dissipation of the magnetic energy (see \\cite{Woods Thompson 2006} for a review). The spectra of the AXP had been modelled well by the compound of a blackbody and a power-law function below $\\sim 10$ keV region. Besides, two temperature blackbody spectrum was also fit well for some AXPs \\cite{Morii et al 2003, Naik et al 2007}. On the other hand, separate hard X-ray emission was discovered for some AXPs above $\\sim 10$ keV \\cite{Kuiper Hermsen Mendez 2004, Kuiper et al 2006}. AXP 1E 1841-045 is located on the center of the supernova remnant (SNR) Kes 73 (G 27.4+0.0) with diameter of about 4$^\\prime$. The kinematic distance toward the SNR was estimated to be between 6 and 7.5 kpc \\cite{Sanbonmatsu Helfand 1992}. The pulse period of 1E 1841-045 was 11.8 s \\cite{Vasisht Gotthelf 1997}. Morii et al. (2003) \\cite{Morii et al 2003} reported that the spectrum was fitted well with the model consisting of the blackbody ($kT = 0.44 \\pm 0.002$ keV) and the power-law function with the hardest photon index among AXP ($\\Gamma = 2.0 \\pm 0.3$), using Chandra data (0.6 -- 7.0 keV). In this analysis, they also showed two blackbody model fits well. Kuiper, Hermsen, \\& Mendez (2004) \\cite{Kuiper Hermsen Mendez 2004} discovered hard X-ray emission up to $\\sim 150$ keV by using RXTE. ", "conclusions": "It is interesting to compare the NPEX parameters between this AXP and accretion-powered pulsars. $\\alpha$ and $-\\beta$ were $\\sim0$ and $\\sim -2$ for accretion-powered pulsars, while those are $3.54 \\pm 0.04$ and $0.92 \\pm 0.06 $ for this AXP (Table \\ref{tab: a}). Both of this AXP are larger than those of accretion powered pulsars by about three. For accretion powered pulsars, power-law component with $\\alpha \\sim 0$ supplies seed photons for Comptonization. On the other hand, for this AXP, the soft power-law component with $\\alpha = 3.54 \\pm 0.04$ would supply seed photons. This steep power-law may be a modified blackbody, which is originated from magnetar's surface and deformed by strong magnetic field of magnetar \\cite{Ozel 2001}. The $\\beta$ parameters can be interpreted as follows. The Comptonization is parameterized by $y$ parameter \\cite{Rybicki Lightman 1979}. For $y >> 1$, $\\beta$ becomes 2, meaning that the spectrum becomes saturated. For accretion-powered pulsars, the emission comes from the accretion column where the plasma density is large, so $y$ become large. On the other hand, the plasma density of this AXP is smaller than those of accretion-powered pulsars. The $y$ parameter become small, then $\\beta$ becomes small. The $y$ must be $\\Gtsim 1$ in this AXP, so that Comptonization process can be effective. This fact suggests that there are corona around magnetosphere of this magnetar \\cite{Beloborodov Thompson 2007}." }, "0801/0801.4795_arXiv.txt": { "abstract": "{The recently-discovered lack of close binaries, among extreme horizontal branch (EHB) stars in Galactic globular clusters, has thus far constituted a major puzzle, in view of the fact that blue subdwarf stars~-- the field counterparts of cluster EHB stars~-- are well-known to present a high binary fraction.} {In this {\\em Letter}, we provide new results that confirm the lack of close EHB binaries in globular clusters, and present a first scenario to explain the difference between field and cluster EHB stars.} {First, in order to confirm that the lack of EHB binaries in globular clusters is a statistically robust result, we undertook a new analysis of 145 horizontal branch stars in NGC\\,6752, out of which forty-one belong to the EHB. To search for radial-velocity variations as a function of time, we repeated high-resolution ($R=18\\,500$) spectroscopy of all stars, four times during a single night of observations.} {We detected a single, hot (25\\,000~K), radial-velocity variable star as a close-binary candidate. From these results, we estimate an upper-limit for the close (period $P \\leq 5$~day) binary fraction $f$ among NGC\\,6752 EHB stars of 16\\% (95\\% confidence level), with the most probable value being $f=4\\%$. Thus our results clearly confirm the lack of close binaries among the hot HB stars in this cluster.} {We suggest that the confirmed discrepancy between the binary fractions for field and cluster EHB stars is the consequence of an $f$-age relation, with close binaries being more likely in the case of younger systems. We analyze theoretical and observational results available in the literature, which support this scenario. If so, an age difference between the EHB progenitors in the field and in clusters, the former being younger (on average) by up to several Gyr, would naturally account for the startling differences in binary fraction between the two populations. } ", "introduction": "\\label{capintro} The presence of a large population of binaries among field B-type subdwarf (sdB) stars, also referred to as Extreme Horizontal Branch (EHB) stars, is well-established in the literature \\citep{Ferguson84,Allard94,Ulla98,Aznar01,Maxted01, Williams01,Reed04,Napiwotzki04}. The measurement of binary fraction varies from one survey to another, probably because of the different survey selection effects. It is generally agreed however, that the binary fraction must be large. Moreover, extensive analysis of orbital parameters \\citep{Moran99,Saffer98,Heber02,MoralesRueda04,MoralesRueda06} reveal that {\\em close} binaries ($P \\leq 10$~d) comprise a major fraction of the field EHB stars, with short-period systems---including either a degenerate or a low-mass main sequence (MS) companion---constituting about half of the entire field sdB population. Indeed, theoretical results indicate that these stars can be naturally explained within the context of binary star evolution \\citep{Han02,Han03}. Moreover, EHB stars of binary origin may also account for the UV flux excess (``UV upturn'') observed in elliptical galaxies \\citep{Han07}, although it should be noted that single-star scenarios have also been suggested \\citep[see][for a recent review]{Catelan07}. One way or another, the link between field sdB stars and binary systems is very well-established, both on observational and theoretical grounds, and close, short-period systems are the most frequently found amongst them. It therefore came as a great surprise that the first radial-velocity (RV) surveys among EHB stars in globular clusters (GCs) revealed a remarkable lack of close binary systems \\citep[see][for a review]{Moni07b}. In this {\\em Letter} we present, in light of new observational results, a first scenario to account for this discrepancy. \\begin{figure*}[t] \\begin{center} \\includegraphics[width=17cm]{res6752.ps} \\caption{Results of our RV variability search. The maximum RV variation (${\\rm RV}_{\\rm max}-{\\rm RV}_{\\rm min}$) observed for each star is plotted at its effective temperature. For the sake of clarity, we plot only the 3$\\sigma$ error bars, which indicate the statistical significance of the observed variations. The data for the only star with a clearly-detected RV variation, is plotted using a larger, solid circle symbol, and thicker lines to indicate its errorbar.} \\label{figresults} \\end{center} \\end{figure*} ", "conclusions": "Using an independent dataset, a sample more than twice as large as in \\citet{Moni06a}, and a resolution in RV variations that is higher by almost a factor of two, for the first time we were able to find a good binary candidate among the EHB stars in \\object{NGC\\,6752}. That notwithstanding, our results confirm that the corresponding (close) binary fraction $f$ is very small in \\object{NGC\\,6752}, with a most likely value of $f = 4\\%$ and an upper limit of $f = 16\\%$ at the 95\\% confidence level. There are hints that a small $f$ is not a peculiarity of \\object{NGC\\,6752}, but could also be a characteristic of other globular clusters \\citep{Moni06b}. This is in sharp contrast with the situation for field sdB stars, where close binaries are at least a factor of ten more frequent, comprising up to 70\\% of the entire sdB population \\citep[see \\S4 in][for a recent review]{Catelan07}. There is however no knowledge about cluster EHB stars in long-period binaries, or with a close low-mass companion, because no survey has investigated their role yet. These kind of systems are known to exist among field sdBs, but are just a minor population, and their presence in GCs would not alleviate the striking contrast with field results. What is the origin of this startling difference between field and cluster EHB stars? We believe that it may not be completely unexpected: there should be a relation between the close binary fraction and the mean age of a sdB population, as a consequence of the different efficiency of binary channels responsible for EHB star formation. Theoretical arguments strongly suggest that sdB stars in close binary systems should have undergone at least one common envelope (CE) phase. Although \\citet{Han02} found an upper limit for the initial mass of the sdB progenitor in this scenario, they also pointed out that within the permitted values a higher initial mass favors the CE channel, and leads to sdB binaries with shorter periods. In fact, a higher mass implies a more tightly-bound envelope, which requires a greater amount of (orbital) energy to be released. On the other hand, \\citet{Han02} explored the stable Roche Lobe Overflow (RLOF) scenario, and found that a higher initial progenitor mass makes it harder for the RLOF to be stable (see their Table~3), because the minimum mass of the companion increases with increasing progenitor mass. For higher values, fewer MS and white dwarf (WD) secondaries are sufficiently massive for the activation of this channel (sdB's with neutron star companions are indeed very rare). The progeny of systems that underwent stable RLOF, that is wide binaries with very long periods, is therefore generated primarily by progenitors of lower initial mass. In light of these results, a relation between $f$ and the mean initial mass of the sdB progenitors could be naturally expected, hence implying a relationship between $f$ and (mean) age. More specifically, one may naturally expect that field sdB's formed from progenitors with a wide spectrum of initial masses (up to about $2\\,M_{\\sun}$), whereas in an old population (such as in globular clusters) only the progeny of less massive stars are currently found on the EHB, those of more massive ones having long evolved away from the He-burning phase. The stable RLOF is an efficient channel for sdB formation in the old case, while the CE one is not---and the CE itself would be released at earlier stages, before the orbits shrink substantially. Therefore, in old populations we should expect to find predominantly wide binaries, or/and single EHB stars formed through other channels \\citep[see][for a recent discussion]{Catelan07}. In fact, WD mergers, the third binary channel studied by \\citet{Han02}, can form (single) sdB stars, and is expected to be particularly efficient for old populations (see below). Moreover, it may be expected to play an important role within the dense environment of a globular cluster, where stellar encounters can harden close binaries \\citep{Heggie75} and thus enhance channel efficiency, as proposed earlier by \\citet{Bailyn89}. However, dynamical effects are not required for the formation of EHB stars \\citep{Whitney98}, as indicated by the lack of radial gradients among the EHB stars in \\object{NGC\\,2808} \\citep{Bedin00} and $\\omega$~Centauri \\citep[\\object{NGC\\,5139};][]{DCruz00}. There are other results in the recent literature that provide support to our proposed scenario. In fact, \\citet{Napiwotzki04} already invoked a possible $f$-age or $f$-metallicity relation to explain their results. Among field sdB stars, they found a close binary fraction lower than in previous surveys, and noted that their sample was on average fainter, possibly including more thick disk or halo (hence older/metal poorer) stars. This suggested difference between the data used has never been fully investigated, and a further check of the populations of these stars would be invaluable. Even stronger support is provided by the recent simulations by \\citet{Han07}, modeling the UV upturn of elliptical galaxies from a binary sdB population. Their Figure~7, though aimed at reproducing the spectral energy distribution of elliptical galaxies, shows how the relative contribution of sdB's formed through different channels evolves with time. One finds that the contribution to the UV flux from RLOF sdB's is essentially constant with time, whereas the one from sdB's that experienced the CE phase first becomes important for a population of age 1.5~Gyr and then slowly fades, being more and more marginal for increasing ages. On the other hand, for populations older than 5~Gyr these authors find that sdB's formed through the merger channel dominate---but these are not binaries any longer. We suggest that these theoretical results implicitly hint at a solution to the heretofore puzzling lack of close EHB binaries in globular clusters, on the one hand, and their large numbers among field stars, on the other. As an alternative to the binary scenario, it is worth noting that many single-star evolutionary channels have been invoked to explain EHB star formation in GCs, including interactions with a close planet \\citep[][see also \\citealt{Silvotti07}]{Soker98}, He mixing driven by internal rotation \\citep{Sweigart79,Sweigart97} or by stellar encounters \\citep{Suda07}, dredge-up induced by H-shell instabilities (\\citealt{vonRudloff88}, but see also \\citealt{Denissenkov03}), close encounters with a central, intermediate-mass black hole \\citep{Miocchi07}, and a sub-population of stars with high helium abundance \\citep[e.g.,][]{DAntona05}. We note that any EHB formation mechanism involving single progenitors should be particularly inefficient among field stars (but possibly not so among GCs), since the EHB progenitors are on average younger, and thus more massive, in the field than in GCs. As a consequence, a much higher amount of RGB mass loss is needed to successfully produce a EHB star in the case of a field progenitor, whereas the ancient GC red giants have a much smaller envelope to be removed before they become EHB stars. This notwithstanding, a small component of single-star progeny may still be needed among field stars to fully explain the available observations \\citep{Lisker05}." }, "0801/0801.2877_arXiv.txt": { "abstract": "Observations show the increase of high-frequency wave power near magnetic network cores and active regions in the solar lower atmosphere. This phenomenon can be explained by the interaction of acoustic waves with a magnetic field. We consider small-scale, bipolar, magnetic field canopy structure near the network cores and active regions overlying field-free cylindrical cavities of the photosphere. Solving the plasma equations we get the analytical dispersion relation of acoustic oscillations in the field-free cavity area. We found that the $m=1$ mode, where $m$ is azimuthal wave number, cannot be trapped under the canopy due to energy leakage upwards. However, higher ($m \\geq 2$) harmonics can be easily trapped leading to the observed acoustic power halos under the canopy. ", "introduction": "Waves play an important role in the dynamics of the solar atmosphere. Observations show an increase of high-frequency power ($\\nu > 5\\;$mHz) in the surroundings of active regions in velocity power maps sometimes called as photospheric power halos \\citep{braun1992,brown1992,hindbro1998,jainheb2002}. The halos were not found in Doppler power maps at lower frequencies ($3\\;$ mHz). Observations also show a lack of power halos in continuum intensity power maps \\citep{hindbro1998,jainheb2002, Mug2005}. On the other hand, recent observations reveal the decrease of the acoustic high frequency power in the chromosphere and its increase in the photosphere near active regions \\citep{Mug2003,Mug2005}. It has also been shown that the quiet-Sun chromospheric magnetic network elements are surrounded by \"magnetic shadows\", which lack the oscillatory power at higher frequency range \\citep{maci2001,kru2001,Vecchio2007}. Therefore, both the photospheric power halos and the chromospheric magnetic shadows probably reflect the same physical process of acoustic wave interaction with overlying magnetic field \\citep[while in subsurface regions the rotational and the meridional non-uniform flows are supposed to have an impact on the formation of the acoustic wave power spectra.][]{sherg2005}. The properties of propagating acoustic waves are closely related to the magnetic field structure. The numerical calculations show that the propagation of acoustic disturbances in the solar atmosphere is strongly determined by the overlying magnetic canopy \\citep{Rosenthal2002,Bogdan2003}. The canopy has been usually modeled with purely horizontal magnetic field \\citep{Evans1990}, but recent high-resolution observations reveal more complex small-scale structure of the field \\citep{dew,centeno2007}. It has been suggested that the magnetic field has small-scale closed loop structure in the vicinity of network cores \\citep{maci2001,schr2003}. The inclined magnetic field may channel low-frequency photospheric oscillations in the chromosphere/corona \\citep{DePont2004}. Here we use a model of small-scale bipolar magnetic canopy near a chromospheric network core and/or an active region. We suggest that granular cells may form field-free cylindrical cavities under the magnetic canopy due to the transport of magnetic flux towards boundaries. These cavities may trap high-frequency acoustic oscillations, while the lower-frequency harmonics may propagate upwards in form of magneto-acoustic waves. Sect. 2 gives the analytical approach, obtained dispersion relation and resulting oscillation spectrum. Sect. 3 includes discussion and comparison of theoretical findings to observations. Sect. 4 briefly summarizes the results. ", "conclusions": "We have studied the spectrum of acoustic oscillations in the cylindrical field-free cavity regions under the small-scale magnetic canopy near the magnetic network cores and active regions. We found that the first harmonic of acoustic oscillations cannot be trapped in the cavity due to the energy radiation by fast magneto-acoustic waves in the canopy. However, the higher ($m \\geq 2$) harmonics can be trapped there, leading to the observed enhancement of high-frequency acoustic power in the photosphere. Future detailed study of the proposed mechanism including gravitational stratification and 3D models is necessary." }, "0801/0801.0043_arXiv.txt": { "abstract": "The Local Supercluster is an ideal laboratory to study distribution of luminous and dark matter in the nearby Universe. The 1100 small groups have been selected using algorithm based on assumption that a total energy of physical pair of galaxies must be negative. The properties of the groups have been considered. ", "introduction": "We used an selection algorithm based on assumption that a total energy of physical pair of galaxies must be negative \\cite{MK00}. Individual mass of the galaxies are estimated from their K-luminosity using appropriate value of M/L. It is chosen to reproduce the well known nearby groups of galaxies, like the Centaurus~A, M~81, M~83 and IC~342. \\begin{figure}[b] \\includegraphics[width=0.32\\textwidth]{lscmass.eps} \\includegraphics[width=0.64\\textwidth]{skymap.eps} \\caption{The mass distribution in the Local Supercluster. The left panel presents distribution of surface mass density in the Local Supercluster plane. The right panel shows an all sky map of surface mass density in the Supergalactic coordinates.} \\label{fig:sky} \\end{figure} ", "conclusions": "" }, "0801/0801.0569_arXiv.txt": { "abstract": "The rapidly oscillating Ap stars are of importance for studying the atmospheric structure of stars where the process of chemical element diffusion is significant. We have performed a survey for rapid oscillations in a sample of 9 luminous Ap stars, selected from their location in the colour-magnitude diagram as more evolved main-sequence Ap stars that are inside the instability strip for rapidly oscillating Ap (roAp) stars. Until recently this region was devoid of stars with observed rapid pulsations. We used the VLT UV-Visual Echelle Spectrograph (UVES) to obtain high time resolution spectroscopy to make the first systematic spectroscopic search for rapid oscillations in this region of the roAp instability strip. We report 9 null-detections with upper limits for radial-velocity amplitudes of $20-65$\\,\\ms\\ and precisions of $\\sigma=7-20$\\,\\ms\\ for combinations of Nd and Pr lines. Cross-correlations confirm these null-results. At least six stars are magnetic and we provide magnetic field measurements for four of them, of which three are newly discovered magnetic stars. It is found that four stars have magnetic fields smaller than $\\sim2$\\,kG, which according to theoretical predictions might be insufficient for suppressing envelope convection around the magnetic poles for more evolved Ap stars. Suppression of convection is expected to be essential for the opacity mechanism acting in the hydrogen ionisation zone to drive the high-overtone roAp pulsations efficiently. Our null-results suggest that the more evolved roAp stars may require particularly strong magnetic fields to pulsate. Three of the studied stars do, however, have magnetic fields stronger than 5\\,kG. ", "introduction": "Why do some stars oscillate while others do not? This question is particularly relevant for the class of rapidly oscillating Ap (roAp) stars for which the discovery of high-frequency oscillations came as a surprise \\citep{kurtz82}. The $\\delta$\\,Scuti pulsations of stars in this area of the colour-magnitude diagram (CMD) were not theoretically predicted to produce the observed high-frequency, high-overtone modes. Subsequently, considerable observational and theoretical efforts have been able to describe most pulsation properties of roAp stars through i) the oblique pulsator model \\citep{kurtz82}, ii) the probable driving mechanism of the pulsations \\citep{balmforthetal01} and iii) their link with the magnetic fields present in Ap stars (see, e.g., \\citealt{cunha06}; \\citealt{saio05}). Ap stars are chemically peculiar stars that range from early-B to early-F spectral types with the majority having detectable magnetic fields with strengths of a few $10^2$ to a few $10^4$\\,G (\\citealt{bychkovetal03}, according to whom $\\sim55$ per cent have mean longitudinal fields above 400\\,G). For the cool Ap stars (also called CP2 stars), abundance anomalies are typically visible in their spectra as abnormally strong absorption lines, particularly of rare-earth elements (REEs). Candidate roAp stars can therefore be selected photometrically due to the influence of these spectral features on narrow-band photometric indices such as those of the Str\\\"omgren filter system (\\citealt{martinez93}, see also Sect.\\,\\ref{sec:selection}). The high-overtone roAp oscillations are thought to be driven by the opacity-mechanism acting in the partial hydrogen ionisation zone \\citep{dziembowskietal96,balmforthetal01,cunha02,saio05}. The frequency range of the excited modes is related to the presence of strong magnetic fields in roAp stars that i) directly stabilise the low-frequency, low-order $\\delta$\\,Scuti oscillations, excited by the $\\kappa$-mechanism acting in the \\ion{He}{ii} partial ionisation zone \\citep{saio05}, and ii) indirectly enhance the driving of high-order oscillations by the $\\kappa$-mechanism acting in the partial hydrogen ionisation zone \\citep{balmforthetal01}. In particular, the stabilisation of the $\\delta$\\,Scuti modes can be explained by the dissipation of slow Alfv\\'en waves \\citep{saio05} and by the suppression of turbulent motion in the outer layers of the magnetic pole regions of roAp stars, which speeds up gravitational settling of helium and drains it from the \\ion{He}{ii} ionisation zone \\citep{theado05}. An implication of the high radial order of the observed modes of roAp stars is that the observable pulsation amplitudes become depth dependent. This pulsational structure provides a unique opportunity for studying the magneto-acoustic structure of the pulsations in 3D because of the element stratification in the atmospheres of these stars. Different elements, such as Fe, Pr, Nd, Eu and H, act as tracers for the pulsation structure due to their different properties regarding formation depth and extent of formation regions. Examples of depth-dependent measurements of pulsations are now numerous, such as \\citet{kurtzetal05, elkinetal2005b, ryabchikovaetal07}. The roAp stars are thus unique objects for studying the interactions among stellar pulsation, magnetic fields and atomic diffusion. The latter is important for, e.g., estimates of globular cluster ages where He settling is significant, pulsation driving mechanisms in sdB and $\\beta$\\,Cephei stars where radiative levitation of Fe is needed for instability, and the Standard Solar Model where both He settling and radiative levitation of some metals are included. Only 37 roAp stars are known at present (\\citealt{kurtzetal06b,tiwarietal07}) despite several searches for rapid pulsation in Ap stars, such as those by \\citet{nelsonetal93, martinezetal94, handleretal99, ashokaetal00, weissetal00, dorokhovaetal05,joshietal06}. The selection of roAp candidates is mostly based on photometric indices pointing to chemical peculiarities in the spectra. There is at present no clear correlation between radial velocity (RV) amplitude and photometric amplitude in roAp stars (see Table~1 of \\citealt{kurtzetal06b}). For a (non-exhaustive) list of spectroscopic studies of roAp stars, see \\citet{kurtzetal06a}. \\citet{cunha02} calculated a theoretical instability strip for roAp stars and compared it to locations of 16 known roAp stars. The region around the terminal age main-sequence is remarkably devoid of pulsators which, as Cunha points out, could be an observational bias. She predicted that more luminous and evolved Ap stars may pulsate with lower frequencies ($0.67 - 0.83$\\,mHz) which makes it harder to detect them in the typically short high-speed observing runs. Note that this frequency range overlaps with the highest $\\delta$\\,Sct frequencies, such as for HD\\,34282 (0.92\\,mHz, \\citealt{amadoetal04}). Alternatively, she proposed that this absence of known evolved roAp stars could reflect a real deficiency in the fraction of pulsators in that region of the HR diagram, resulting from the fact that the magnetic field is less likely to suppress envelope convection in more evolved stars. The photometric survey by \\citet{martinezetal94} was indeed less sensitive to low-amplitude roAp periods longer than 15\\,min and got a null-result for HD\\,116114 for which a 21-min oscillation was recently detected spectroscopically by \\citet{elkinetal05}. A dedicated survey of luminous Ap stars therefore seems pertinent. Ap stars inside the roAp instability strip that do not exhibit any detectable variability are called non-oscillating Ap (noAp) stars. As seen in, e.g., the astrometric HR-diagram by \\citeauthor{hubrigetal05} (\\citeyear{hubrigetal05}, their figure 2) for roAp and Ap stars, the apparent noAp stars occupy essentially the same regions as the roAp stars. However, the noAp stars appear to be systematically more evolved than the roAp stars \\citep{north97,handleretal99,hubrig00}. To fully understand the mechanism responsible for driving oscillations in roAp stars, it is essential to confirm and understand why so many stars with similar characteristics are apparently stable against pulsations. The apparent absence of oscillations in noAp stars does not, however, necessarily mean that oscillations are suppressed. For instance, detection of oscillations may depend on lifetimes of excited modes (may be days for some roAp stars, \\citealt{handler04}), on beating between multiple modes, on the geometry of the magnetic field's orientation and the aspect in which the stellar surface is observed together with the surface distribution of the chemical elements used to detect the oscillations. Further, the signal-to-noise ratio ($S/N$) of the obtained photometry or spectroscopy may simply be too low to detect the typically low-amplitude roAp pulsations. Finding high precision null-results for a statistically significant sample of roAp candidates would be a safe basis for concluding whether noAp stars exist or whether the current lack of known luminous roAp stars is an observational bias. From a theoretical point of view, stars located inside the theoretical instability strip may be noAp stars if the magnetic field is too weak to effectively suppress convection in the stellar envelope. A good test sample should, therefore, also include stars with strong magnetic fields. To try to answer the question of whether the absence of observed variability among the luminous Ap stars is an observational selection effect only, or in fact due to intrinsic properties, we have studied 9 such stars at high radial-velocity precision. Our immediate goal with UVES was to search these luminous Ap stars for roAp oscillations at high radial velocity precision. Table\\,\\ref{tab:obslog} lists the known properties of the 9 targets and gives an observing log for the collected spectra. We find no pulsation in any of the 9 stars; Tables\\,\\ref{tab:cogpow} and \\ref{tab:ccpow} below present the null-results of the frequency analyses. In the following sections, we describe selection and observation of the targets along with the data reduction in Sect.\\,2, then follows the data analyses including estimation of physical parameters and the radial-velocity analysis in Sect.\\,3. We show that our analyses reach the same precision as other radial-velocity studies of roAp stars in the literature and we then give the results for each target on a star-by-star basis. Finally we discuss the null-results in Sect.\\,4. \\begin{table*} \\centering \\begin{minipage}{140mm} \\caption{\\label{tab:obslog}Observing log indicating number of observations and the mid-series heliocentric Julian Date for each series of spectra. Exposure times were 40\\,s, except for the fainter stars HD\\,110072 and HD\\,170565 where 80\\,s was used; readout and setup times were $24-27$\\,s. The last two columns indicate $S/N$ measured in the continuum based on the standard deviation in the residual flux from two consecutive spectra. {\\it Lower} and {\\it Upper} respectively refer to wavelengths below or above the gap between the two CCDs at 6000\\,\\AA, while the ranges in parentheses correspond to the $S/N$ range for all spectra in that series. The $S/N$ estimates assume random noise only.} \\begin{tabular}{@{}lrccccll@{}} \\hline Star &\\#spec& $\\alpha_{2000.0}$&$\\delta_{2000.0}$ &HJD\\,obs&Time&S/N Lower& $S/N$ Upper \\\\ & & (h:m:s) & (\\degr:\\,\\arcmin\\,:\\,\\arcsec\\,) &(d)&(h)& & \\\\ \\hline HD\\,107107 & 111 & 12:19:04.8 & -40:09:47 & 2453509.504 &1.94 & 79\\phantom{1} (77--82) & 62\\phantom{1} (56--77) \\\\ HD\\,110072 & 69 & 12:39:50.2 & -34:22:30 & 2453509.591 &2.06 & 50\\phantom{1} (43--55) & 38\\phantom{1} (32--49) \\\\ HD\\,131750 & 111 & 14:56:20.7 & -30:52:36 & 2453509.681 &1.98 & 80\\phantom{1} (74--87) & 66\\phantom{1} (57--74) \\\\ HD\\,132322 & 111 & 15:01:36.0 & -63:55:35 & 2453510.519 &2.03 &151 (147--155) & 141 (132--152)\\\\ HD\\,151301 & 111 & 16:49:28.3 & -54:26:47 & 2453510.715 &1.95 & 82\\phantom{1} (76--84) & 67\\phantom{1} (60--79) \\\\ HD\\,170565 & 85 & 18:30:08.2 & -02:35:26 & 2453509.778 &2.50 & 89\\phantom{1} (83--94) & 77\\phantom{1} (69--86) \\\\ HD\\,197417 & 125 & 20:48:48.1 & -72:12:44 & 2453509.879 &2.20 &107 (94--113) & 93\\phantom{1} (79--110)\\\\ HD\\,204367 & 111 & 21:28:41.2 & -25:38:39 & 2453510.798 &1.96 &113 (108--120) & 101 (93--115)\\\\ HD\\,208217 & 138 & 21:56:56.4 & -61:50:44 & 2453510.896 &2.52 &156 (142--165) & 148 (128--159)\\\\ \\hline \\hline \\end{tabular} \\end{minipage} \\end{table*} ", "conclusions": "The class of roAp stars is notoriously difficult to supplement with new members, as demonstrated by several photometric and spectroscopic studies before this one. But motivated by the recent discovery of the luminous roAp star HD\\,116114 \\citep{elkinetal05}, we have spectroscopically tested a sample of 9 luminous Ap stars for rapid pulsations. Using lines known to show pulsations in roAp stars, we reach typical upper amplitude limits in radial velocity of: $40-75$\\,\\ms\\ ($\\sigma=13-24$\\,\\ms) for the line core of \\halpha, $20-65$\\,\\ms\\ ($\\sigma=7-20$\\,\\ms) when combining all measured Nd and Pr lines, and $20-40$\\,\\ms\\ ($\\sigma=7-11$\\,\\ms) when combining all measured lines. With cross-correlations, using large wavelength regions, we typically reach upper amplitude limits of $4-10$\\,\\ms\\ ($\\sigma=1-4$\\,\\ms). In spite of a clear theoretical prediction \\citep{cunha02} and empirical (HD\\,116114) evidence for roAp pulsations in this part of the Hertzsprung--Russell diagram, we end up with 9 null-results, or noAp stars. A number of questions are therefore pertinent to discuss. \\vspace{2mm} \\noindent{\\em How well does our test sample resemble known roAp stars?} \\vspace{1mm} \\noindent All studied stars have strong REE lines, except for HD\\,204367 and possibly also HD\\,197417. They have the core-wing anomaly typical for roAp stars, and {\\it Hipparcos} luminosities with our temperature estimates place them inside the predicted roAp instability strip. Several appear to have spotted surface distributions of REEs (such as HD\\,170565, HD\\,151301, HD\\,132322 and perhaps also HD\\,197417) which is typically associated with the strong magnetic fields common in known roAp stars. Indeed most of the stars are magnetic, and cover a range in magnetic field strengths of $0.4-8.0$\\,kG, comparable to that of known roAp stars \\citep{kurtzetal06b}. In the case of the sharp-lined HD\\,110072, we compared its spectrum in detail with those of two known roAp stars, and found remarkable similarities for the REEs. However, HD\\,110072 and HD\\,208217 are double-lined and single-lined binaries, respectively, which might indirectly influence their stability to high-frequency pulsations by reducing the magnetic field intensity (see \\citealt{cunha02} and references therein). Still, the orbits are probably too wide in both cases for tidal interaction to occur and HD\\,208217 has a known strong magnetic field ($\\left=8$ kG). The known roAp stars have cases of wide binaries, such as $\\beta\\,$CrB (spectroscopic binary), HR\\,3831, $\\alpha$\\,Cir, $\\gamma$\\,Equ and HD\\,99563 (visual binaries). In these regards, the studied sample has the characteristics of roAp stars. \\vspace{2mm} \\noindent{\\em Could pulsations have been overlooked?} \\vspace{1mm} \\noindent\\citet{kurtzetal06b} published radial velocity amplitudes for the \\halpha\\ cores of 16 roAp stars. Of these, only 3 have amplitudes below 75\\,\\ms\\ (HD\\,116114, HD\\,154708 and HD\\,166473, of which two have amplitudes above 3\\,$\\sigma$), while the rest range from $148-2528$\\,\\ms. Further, radial velocity series for Pr and Nd line measurements in UVES spectra of these 16 roAp stars (Kurtz et al., partly unpublished), show that about 75 per cent of the measured and significant (3\\,$\\sigma$) amplitudes are within the range $350-1600$\\,\\ms. Five of these stars have very small Nd and Pr amplitudes ($60-90$\\,\\ms): HD\\,166473, HD\\,116114, $\\beta$\\,CrB, 33\\,Lib and HD\\,154708. In such difficult cases, other lines or combinations of several lines makes detection of pulsations possible. We successfully tested our procedures on the two latter roAp stars in Sect.\\,\\ref{sect-rvshift}, and also used combinations of several lines, including of different elements. More of the other 19 known roAp stars have low amplitudes, such as 10\\,Aql, but our tests and analyses show that we reach these amplitude levels and should have detected such rapid pulsations if present in the studied sample. A complication for our analyses is typical \\vsini\\,$\\sim10-30$\\kms combined with double lines due to either spots on the stellar surface and/or magnetic splitting, which results in considerable line blending and makes line identification and analysis more difficult. However, our radial-velocity analysis was based partly on cross-correlations that are more robust than line measurements (as shown by our tests for two roAp stars) and this method similarly results in flat amplitude spectra that exclude rapid oscillations to relatively small roAp-amplitude levels. It also seems improbable that, e.g., unfavourable viewing angles of the global pulsations or short mode lifetimes could explain a momentary lapse of detectable pulsation amplitudes simultaneously in all nine stars. In fact, we know independently that HD\\,208217 was observed near its magnetic negative extremum where roAp pulsations are expected to have maximum amplitudes. Future surveys like this one may benefit from being repeated at different rotation phases. We also note that the near-normal REE abundances of HD\\,197417 and HD\\,204367 reduce the probability that they are roAp stars, given the strong peculiarity of all known roAp stars. \\vspace{2mm} \\noindent {\\em Do these stars really not pulsate?} \\vspace{1mm} \\noindent In pulsators such as roAp stars oscillations are intrinsically unstable. Their excitation depends on the balance between the driving and damping of the oscillations over each pulsation cycle. In roAp stars this balance is thought to be particularly delicate. On one hand, the amount of energy input through the opacity mechanism acting on the hydrogen ionisation region depends strongly on the interaction between the magnetic field and envelope convection, being maximal in the regions where envelope convection is suppressed \\citep{balmforthetal01,cunha02}. On the other hand, the direct effect of the magnetic field on pulsations can introduce significant energy losses, through slow Alfv\\' en waves in the interior and through acoustic waves in the atmosphere, both resulting from mode conversion in the magnetic boundary layer \\citep{cunha00,saio05}. Due to this delicate balance, it is not too surprising that roAp and noAp stars occupy the same locus in the HR diagram. Despite the developments in theoretical studies of linear non-adiabatic pulsations in models of roAp stars, we still lack a theoretical study that takes into account all these phenomena simultaneously and, thus, cannot firmly predict the conditions under which pulsations should be expected in roAp stars. In fact, both studies of \\cite{cunha02} and \\cite{saio05} considered the extreme case in which envelope convection is fully suppressed. Moreover, the first of these studies did not consider the direct effect of the magnetic field on pulsations and neglected the energy losses as a result of mode conversion, and the second study, while considering mode conversion, assumed the waves are fully reflected at the surface, hence neglecting energy losses through acoustic running waves in the atmosphere. As discussed by \\cite{cunha02}, the condition for suppression of envelope convection, which seems necessary to make the high frequency modes unstable, is in principle harder to fulfil in evolved stars due to the increase with age of the absolute value of the buoyancy frequency in the region where hydrogen is ionised. Hence, it is likely that in evolved stars oscillations are excited only if the magnetic field is relatively strong. Unfortunately, the complexity of the interaction between magnetic field and convection makes it impossible to derive a global convective stability criterion, even if local criteria for convective stability may be established \\citep{Gough66,Moss69} \\citep[see also][for an extensive discussion on this subject]{theado05}. Thus, the magnetic intensity needed to suppress convection at a given age, for a given mass, is very hard to establish. The more evolved roAp star HD\\,116114, in which a relatively low frequency oscillation was found well in agreement with theoretical predictions, has a magnetic field modulus of $\\approx 6$~kG. In contrast with this, most stars in our sample have estimated mean magnetic field moduli around or below 2~kG. The clear exceptions are HD\\,107107, HD\\,131750 and HD\\,208217. Of these three, the latter is clearly an important test case to check the theoretical predictions. It has the strongest confirmed magnetic field in our sample and is strongly peculiar. However, we observed HD\\,208217 when one of its magnetic poles were almost visible, so pulsation should have been near its maximum amplitude. From the observational point of view, one way to investigate the conditions under which roAp star oscillations are excited, and thus test theoretical models, is by identifying systematic differences between roAp and noAp stars. This study would have been able to detect pulsations in all the known roAp stars, and any missed rapid pulsations must have amplitudes lower than these. Hence we conclude that based on the obtained data, most stars in our sample are indeed noAp stars, and also excellent roAp candidates. Despite this conclusion, it is premature to state that we can confirm the evidence that noAp stars are in average more luminous and more evolved than roAp stars, as indicated by earlier studies based on photometric surveys for pulsations in roAp candidates \\citep{north97,handleretal99,hubrig00}. To conclude that, we would also need to search for rapid pulsations in a control group of less evolved roAp stars with lower luminosity and compare the frequency of null-results in the two cases. Such a survey has been started and results are expected in the near future. The next step is to analyse the noAp stellar atmospheres in detail, taking temperature gradients and the abundance stratification into account. New spectra are needed of HD\\,132322 to clarify the origin of its double line structures, and of HD\\,110072 to verify its secondary spectrum and to test for the `spurious' absorption lines in the recent FEROS spectrum. Polarimetry of HD\\,110072 and HD\\,204367 is needed to confirm the detected magnetic fields." }, "0801/0801.2270_arXiv.txt": { "abstract": "{} {The present work provides the results of the first six years of operation of the systematic night-sky monitoring at ESO-Paranal (Chile).} {The $UBVRI$ night-sky brightness was estimated on about 10,000 VLT-FORS1 archival images, obtained on more than 650 separate nights, distributed over 6 years and covering the descent from maximum to minimum of sunspot cycle n. 23. Additionally, a set of about 1,000 low resolution, optical night-sky spectra have been extracted and analyzed.} {The unprecedented database discussed in this paper has led to the detection of a clear seasonal variation of the broad band night sky brightness in the $VRI$ passbands, similar to the well known semi-annual oscillation of the Na~I~D doublet. The spectroscopic data demonstrate that this seasonality is common to all spectral features, with the remarkable exception of the OH rotational-vibrational bands. A clear dependency on the solar activity is detected in all passbands and it is particularly pronounced in the $U$ band, where the sky brightness decreased by $\\sim$0.6 mag arcsec$^{-2}$ from maximum to minimum of solar cycle n.~23. No correlation is found between solar activity and the intensity of the Na~I~D doublet and the OH bands. A strong correlation between the intensity of N~I 5200\\AA\\/ and [OI]6300,6364\\AA\\/ is reported here for the first time. The paper addresses also the determination of the correlation timescales with solar activity and the possible connection with the flux of charged particles emitted by the Sun.} {} ", "introduction": " ", "conclusions": "" }, "0801/0801.2100_arXiv.txt": { "abstract": "We investigate the evolution of accretion luminosity $L_{\\rm acc}$ and stellar luminosity ${L_\\ast}$ in pre-mainsequence stars. We make the assumption that when the star appears as a Class II object, the major phase of accretion is long past, and the accretion disc has entered its asymptotic phase. We use an approximate stellar evolution scheme for accreting pre-mainsequence stars based on Hartmann, Cassen \\& Kenyon, 1997. We show that the observed range of values $k = L_{\\rm acc}/L_\\ast$ between 0.01 and 1 can be reproduced if the values of the disc mass fraction $M_{\\rm disc}/M_*$ at the start of the T~Tauri phase lie in the range 0.01 -- 0.2, independent of stellar mass. We also show that the observed upper bound of $L_{\\rm acc} \\sim L_\\ast$ is a generic feature of such disc accretion. We conclude that as long as the data uniformly fills the region between this upper bound and observational detection thresholds, then the degeneracies between age, mass and accretion history severely limit the use of this data for constraining possible scalings between disc properties and stellar mass. ", "introduction": "The claimed relationship between accretion rate, $\\dot{M}$, and mass, $M_*$, in pre-mainsequence stars (e.g. Natta, Testi \\& Randich 2006 and references therein) originates from the more direct observational result that in Class II pre-mainsequence stars the accretion luminosity, $L_{\\rm acc}$, is similarly found to correlate with stellar luminosity, $L_\\ast$. The accretion luminosity is deduced from the luminosity in the emission lines of hydrogen (in Natta et al., 2006, these are Pa$\\beta$ and Br$\\gamma$), together with modeling of the emission process (Natta et al. 2004; Calvet et al. 2004). The modeling appears to be relatively robust. In order to deduce the quantities $\\dot{M}$ and $M_*$ from the quantities $L_{\\rm acc}$ and $L_\\ast$, it is necessary to be able to deduce the stellar properties. This is done by placing the object in a Hertzsprung-Russell diagram and making use of theoretical pre-mainsequence tracks for non-accreting stars (e.g. D'Antona \\& Mazzitelli 1997). The result of this exercise indicated a correlation between $\\dot{M}$ and $M_*$ of the form $\\dot{M} \\propto M_*^{\\alpha}$, where $\\alpha$ is close to $2$. The simplest theoretical expectation (in the case of disc accretion where neither the fraction of material in the disc nor the disc's viscous timescale scale systematically with stellar mass) is instead that $\\dot{M} \\propto M_*$. The claimed steeper than linear relationship thus motivated a variety of theoretical ideas about specific scalings of disc parameters with stellar mass (Alexander \\& Armitage 2006, Dullemond et al 2006). \\begin{figure} \\resizebox{\\colwidth}{!}{\\includegraphics[angle=0]{fig1_till.eps}} \\caption{The distribution of Classical T Tauri stars in Ophiuchus in the $L_\\ast$, $L_{\\rm acc}$ plane. Detections (filled squares) and upper limits (crosses) deduced from emission line data. Adapted from Natta et al 2006.} \\label{hereifany2} \\end{figure} Clarke \\& Pringle (2006) however suggested that the claimed correlation between $\\dot{M}$ and $M_*$ could be understood as follows. They noted that the distribution of datapoints in the plane of $L_\\ast$ versus $L_{\\rm acc}$ (Figure 1) more or less fills a region that is bounded, at high $L_{\\rm acc}$ by the condition $k = L_{\\rm acc}/L_\\ast \\sim 1$ and, at low $L_{\\rm acc}$, by observational detection thresholds, which roughly follow a relation of the form $L_{\\rm acc} \\propto L_\\ast^{1.6}$. They noted that when this relation between $L_{\\rm acc}$ and $ L_\\ast$ is combined with specific assumptions about the relationship between $L_\\ast$ and $M_*$, and between $L_\\ast$ and stellar radius $R_\\ast$ (based on placing the stars on pre-mainsequence tracks at a particular age, or narrow spread of ages) then the resulting relationship between $\\dot{M}$ and $M_*$ is indeed $\\dot{M} \\propto M_*^2$. They thus claimed that it is currently impossible to reject the possibility that the claimed, steeper than linear, relationship is an artefact of detection biases. Clarke \\& Pringle (2006) also pointed out that the distribution of detections in the $L_{\\rm acc}$ -- $L_\\ast$ plane may nevertheless be used to tell us something about conventional accretion disc evolutionary models. In this paper we explore the extent to which this might be achieved. The first question which needs to be addressed is why there is a spread of values of $L_{\\rm acc}$ at a given value of $L_\\ast$. Secondly we need to understand {\\it why} the upper locus of detections corresponds to $L_{\\rm acc} \\sim L_\\ast$, since there seems no obvious reason why this should correspond to a detection threshold or selection effect. We begin here (Section 2) by setting out a simple argument why -- as a generic property of accretion from a disc with mass much less than the stellar mass -- one would expect the distribution to be bounded by the condition $L_{\\rm acc} \\sim L_\\ast$. In Section 3 we use the ideas of Hartmann, Cassen \\& Kenyon (1997) to develop an approximate evolutionary code for accreting pre-mainsequence stars and demonstrate its applicability by comparison with the tracks of pre-mainsequence stars computed by Tout, Livio \\& Bonnell (1999). In Section 4 we apply the code to pre-mainsequence stars accreting from an accretion disc with a declining accretion rate and investigate the range of disc parameters required to obtain the observed range of $L_{\\rm acc}/L_\\ast$. We draw our conclusions in Section 5. \\section {The maximum ratio of accretion luminosity to stellar luminosity} From a theoretical point of view there are three relevant timescales in the case of star gaining mass by accretion from a disc. The first is \\begin{equation} t_M = M_*/\\dot{M}, \\end{equation} which is the current timescale on which the stellar mass is increasing. The second is \\begin{equation} t_{\\rm disc} = M_{\\rm disc}/\\dot{M}, \\end{equation} which is the current timescale on which the disc mass is decreasing. The third is the Kelvin-Helmholtz timescale \\begin{equation} t_{\\rm KH} = \\frac{GM_*^2}{R_\\ast L_\\ast}, \\end{equation} which is the timescale on which the star can radiate its thermal energy. It is also the timescale on which a star can come into thermal equilibrium. On the pre-main sequence, when the luminosity of the star is mainly due to its contraction under gravity, $t_{\\rm KH}$ is also the evolutionary timescale. We see immediately (since $L_{\\rm acc} \\sim G M_* \\dot M/R$) that \\begin{equation} \\frac{L_{\\rm acc}}{L_\\ast} = \\frac{t_{\\rm KH}}{t_M}. \\end{equation} The pre-mainsequence stars in which we are interested, Class~II (the Classical T~Tauri stars), all have the following properties: (i) They have passed through the major phase of mass accretion which occurs during the embedded states Class 0 and Class I. This implies typically that they have been evolving for some time with their own gravitational energy as the major energy source. Deuterium burning interrupts this briefly but only delays contraction by a factor of 2 or so. Thus for these stars we may expect that they have an age $T \\approx t_{\\rm KH}$ roughly. (ii) The masses of their accretion discs are now small, $M_{\\rm disc} < M_*$. Further, accretion disc models at such late stages -- i.e. the so called asymptotic stage when most of the mass that was originally contained in them has been accreted on to the central object -- exhibit a power law decline in $\\dot {M}$ with time and thus have the generic property that the age of the disc is about $t_{\\rm disc}$. If most of the star's lifetime has been spent accreting in this asymptotic regime, we may roughly equate the age of the disc and the age of the star and write $T \\approx t_{\\rm disc}$. Given this, we see immediately that, for these stars, we expect $t_M = M_*/\\dot{M} > M_{\\rm disc}/\\dot{M} = t_{\\rm disc} \\approx t_{\\rm KH}$ and therefore that $L_{\\rm acc} < L_\\ast$. In the following section we test this argument by evolving a suite of model star-disc systems in the $L_\\ast - L_{\\rm acc}$ plane. ", "conclusions": "We have shown, using simplified stellar evolution calculations for stars subject to a time dependent accretion history, that a plausible morphology of the $L_* - L_{\\rm acc}$ plane for T~Tauri stars has the following features, (i) an upper locus corresponding to $L_* \\approx L_{\\rm acc}$ and (ii) diagonal tracks for falling accretion rates as stars descend Hayashi tracks. We have argued that (i) is a generic feature of stars which are accreting from a disc reservoir with mass less than the central star coupled with the assumption that the mass depletion timescale of the disc is comparable to the age of the star. \\footnote {We may in fact argue that the existence of such an upper limit in the data is a good signature of temporally declining accretion through a disc, since if Classical T Tauri stars instead accreted external gas at a constant rate (e.g. Padoan et al 2005), no such feature would be apparent in the data}. The first is necessary for stars to be classified as Classical T~Tauri stars while the second results from the gradual decay of the accretion rate. We show that the slope of of the diagonal tracks (ii) is related to the index of the power law decline of accretion rate with time at late times. This itself depends on the scaling of viscosity with radius in the disc. For reference, if $\\nu \\propto R$, then $L_{\\rm acc} \\propto L_\\ast^{1.7}$ along such tracks, whereas if $\\nu \\propto R^{1.5}$ then $L_{\\rm acc} \\propto L_\\ast^{2.4}$. We however argue that, as long as observational datapoints appear to fill the whole region of the $L_\\ast - L_{\\rm acc}$ plane that is bounded by the upper limit (i) and luminosity dependent sensitivity thresholds, it will remain impossible to deduce what viscosity law or what dependence of disc parameters on stellar mass will be required in order to populate this region. This is because of the severe degeneracies that exist between stellar mass, age and accretion history (our results show that in order to populate the desired region, it is necessary that there is, at the least, a spread in both stellar mass and initial disc properties). It is not, however, necessary to posit any particular scaling of disc properties with stellar mass in order to populate the plane, in contrast to the hypotheses of Alexander \\& Armitage 2006 and Dullemond et al 2006. Likewise, although, as noted above, the trajectories followed by individual stars in this plane are a sensitive diagnostic of the disc viscosity law, this information is washed out in the case of an ensemble of stars which simply fill this plane. We therefore conclude that such data could only yield useful insights if regions of the plane turned out to be unpopulated (or very sparsely populated). There is a shadow of a suggestion in Figure 1 that at low $L_\\ast$ the data falls further below the locus $L_{\\rm acc} = L_\\ast$ than at higher luminosities although, given the sample size, this difference is not statistically significant (Clarke \\& Pringle 2006). Any such edge in the distribution could in principle re-open the possibility of constraining disc models. We however emphasise that this low luminosity regime can in any case not be explored by the simple evolutionary models that we employ here, and that -- if it becomes necessary to investigate this regime -- it will be essential to use full stellar evolution calculations. Finally, we note that we have chosen to relate our study to observational data in the $L_{\\rm acc} -L_*$ plane because the ratio of these quantities is straightforwardly derivable from emission line equivalent widths. Thus this comparison is more direct than the alternative, i.e. the comparison with observational data that has been transformed into the plane of $\\dot M$ versus $M_*$ via the use of isochrone fitting and the application of theoretical mass-radius relations. It is nevertheless worth noting that the upper locus of $L_{acc} \\sim L_*$ corresponds to a line with $\\dot M \\propto M_*$. Although it is often taken as self-evident that this is the expected ratio (in the case that the maximum initial ratio of disc mass to stellar mass is independent of stellar mass), this actually only corresponds to $\\dot M \\propto M_*$ in the case that the disc's initial viscous time is independent of stellar mass. Given the possible variety of viscosity values and radii of protostellar discs it is not at all evident that these should conspire together to give the same viscous timescale -- until, that is, that one realises that all discs in the asymptotic regime have viscous timescale which is equal to the disc age. The conclusion that the upper locus should follow the relation $\\dot M \\propto M_*$ is thus a very general one. In summary, we conclude, as in the study of Clarke \\& Pringle (2006), that the observed population of the $L_{\\rm acc} - L_*$ plane, and the consequent relationship between $\\dot {M}$ and $M_*$, is strongly driven by luminosity dependent sensitivity thresholds and an upper locus at $L_{\\rm acc} \\approx L_\\ast$. The main new ingredient injected by our numerical calculations and plausibility arguments is that we now understand that this observed upper locus can be understood as a consequence of plausible accretion histories, given the current properties of the relevant pre-mainsequence stars. In addition, our computations indicate the range of initial disc properties that are required in order to populate the region of parameter space occupied by observational datapoints: we are able to reproduce the observed range of values of $L_{\\rm acc}/L_\\ast$ at a given value of $L_\\ast$ by assuming once the T~Tauri phase has been reached, the disc mass fractions lie roughly in the range $0.01 \\le M_{\\rm disc}/M_* \\le 0.2$, independent of stellar mass." }, "0801/0801.0333_arXiv.txt": { "abstract": "We consider a gauge inflation model in the simplest orbifold $M_4 \\times S^1/\\mathbb{Z}_2$ with the minimal non-Abelian $SU(2)$ hidden sector gauge symmetry. The inflaton potential is fully radiatively generated solely by gauge self-interactions. Following the virtue of gauge inflation idea, the inflaton, a part of the five dimensional gauge boson, is automatically protected by the gauge symmetry and its potential is stable against quantum corrections. We show that the model perfectly fits the recent cosmological observations, including the recent WMAP 5-year data, in a wide range of the model parameters. In the perturbative regime of gauge interactions ($\\gfd \\lesssim 1/(2\\pi R\\mpl)$) with the moderately compactified radius ($10 \\lesssim R\\mpl \\lesssim 100$) the anticipated magnitude of the curvature perturbation power spectrum and the value of the corresponding spectral index are in perfect agreement with the recent observations. The model also predicts a large fraction of the gravitational waves, negligible non-Gaussianity, and high enough reheating temperature. ", "introduction": "It is now widely accepted that an early period of accelerated expansion of the universe, or inflation~\\cite{inf}, can resolve many cosmological problems such as horizon problem and can provide the desired initial conditions for the subsequent hot big bang evolution of the observed universe~\\cite{books}. Many observational facts such as the flatness of the universe and the isotropy of the cosmic microwave background (CMB) are natural consequences of inflation, and hence they strongly support the existence of such a period of acceleration in the very early universe. In particle physics point of view, inflation occurs when one or more scalar fields, the inflaton fields, dominate the energy density of the universe with their potential being overwhelming~\\cite{Lyth:1998xn}. Under such a condition, dubbed slow-roll condition, curvature perturbation $\\mathcal{R}$ is produced which is nearly scale invariant and is heavily constrained by the measurements of the anisotropies of the CMB and the observations of the large scale structure~\\cite{obs}, including the recent Wilkinson Microwave Anisotropy Probe (WMAP) 5-year data set~\\cite{Komatsu:2008hk}. The slow-roll condition says that the inflaton potential should be very flat, i.e. the effective mass of the inflaton should be very small compared with the inflationary Hubble parameter. This is, however, quite difficult to be achieved: for example, in supergravity, any generic scalar field is expected to have an effective mass of $\\mathcal{O}(H)$~\\cite{sugrainf}, which completely spoils the desired slow-roll condition. It is thus natural to consider some symmetry principle which protects the small inflaton mass from large radiative corrections or supergravity effects. Shift symmetry, under which the field is invariant with respect to the transformation \\begin{equation} \\phi \\to \\phi + a \\, , \\end{equation} with $a$ being an arbitrary constant, is one of such symmetries and the corresponding fields remain completely massless, i.e. their potential is exactly flat as long as shift symmetry is unbroken. When the symmetry is explicitly broken, they become pseudo Nambu-Goldstone bosons (pNGBs) and the potential acquires a tilt depending on the symmetry breaking scale $f$ and is of the form \\begin{equation}\\label{naturalinf_V} V(\\phi) = \\Lambda^4 \\left[ 1 \\pm \\cos\\left( \\frac{\\phi}{f} \\right) \\right] \\, . \\end{equation} The model of natural inflation~\\cite{naturalinf} makes use of such a pNGB with large $f$, which renders the potential very flat. This largeness, however, requires $f \\gg \\mpl \\equiv G^{-1/2} \\sim 10^{19} \\mathrm{GeV}$ and hence the valid region of successful natural inflation lies in the regime where we completely lose theoretical control. This has been regarded as a significant drawback of natural inflation. An idea to evade this problem was suggested in Ref.~\\cite{Arkani-Hamed:2003wu}\\footnote{The idea of identifying the higher dimensional gauge field as the inflaton is also suggested in Ref.~\\cite{Kaplan:2003aj}.}. In this scenario, called gauge inflation, the inflaton field is essentially coming from a part of the fifth component of the gauge boson in five dimensional bulk $A_5$, by which the gauge invariant Wilson line is defined as \\begin{equation} e^{i\\theta} \\equiv \\exp \\left[i g \\oint A_5 dy \\right] \\, , \\end{equation} where $g$ is the gauge coupling constant. By the one-loop interactions with the charged particles, the potential of the canonically normalized field $\\phi \\equiv f\\theta$ has basically the same form as Eq.~(\\ref{naturalinf_V}) while $f=1/(g_4 L)$ has extra dimensional nature with $L$ being the size of the fifth dimension. Hence the potential can be trusted even when $f$ is larger than $\\mpl$ in the perturbative regime of gauge interaction, $g_4 \\lesssim 1/(\\mpl L)$. An interesting possibility to unify the grand unification theory (GUT) and cosmic inflation in the framework of the gauge inflation was suggested in Ref.~\\cite{Park:2007sp} by one of the authors of the present paper: starting from the higher dimensional $SU(5)$ GUT, the standard model is recovered by orbifold projection by $S^1/{\\mathbb{Z}_2}$ and a massless scalar boson from the fifth component of the gauge boson is shown to play the role of the inflaton with a fully radiatively generated flat potential. In this paper, we take a different approach. Here we try to build the minimal model of gauge inflation using the simplest orbifold $S^1/\\mathbb{Z}_2$ and the minimal non-Abelian gauge group $SU(2)$ for the hidden sector. The obvious advantages of this approach is two-fold. First, by taking the hidden sector gauge interaction, the theory is less constrained than the visible sector gauge theory, e.g. GUT~\\cite{Park:2007sp}. This is important since the required coupling constant is typically very small, $g_4 \\sim 10^{-2}$, so it is hard to reconcile this smallness with the coupling constant unification. Second, the minimality can apply to the particle contents. Different from the Abelian case, the non-Abelian gauge theory accommodates the gauge self-interactions by which the potential for the inflaton ($\\sim A_5$) can be generated even without any additional charged matter fields. So the theory remains minimal in particle contents and its predictions are quite robust and free from parameters such as the masses and other quantum numbers of newly introduced matter fields. Combining the minimal choice of the orbifold and the gauge group, no exotic particles appear in the theory. The model is clean. The structure of the paper is as follows. In the next section, we give the detail of the higher dimensional gauge theory and outline the subsequent cosmological scenario. Section~\\ref{sec_cos} is devoted to the cosmological evolution of the model and we analytically calculate the observable quantities produced during inflation and study their relations to the parameters of the model. We also address the issue of reheating and estimate the reheating temperature $T_\\mathrm{RH}$. In Section~\\ref{sec_con} we then conclude. ", "conclusions": "\\label{sec_con} In this paper, we have presented a cosmological scenario from the hidden sector $SU(2)$ gauge symmetry in the five dimensional orbifold $M_4\\times S^1/\\mathbb{Z}_2$. The model is minimal in several aspects: the minimal non-Abelian gauge group and the minimal orbifold compactification with the minimal number of extra dimensions. Thanks to the non-Abelian nature, the bulk gauge boson, the fifth component $A_5$ in particular, could have a one-loop induced effective potential without introducing any exotic field in the model. This makes sure the minimality of the model. The advantage of this minimal setup is as follows: the inflaton field is a built-in ingredient of the theory and is automatically free from quantum gravitational effects because of its higher dimensional locality and the gauge symmetry. Fully radiatively generated one-loop potential is naturally able to support a long enough period of slow-roll inflation provided that the theory is weakly coupled, i.e. $\\gfd \\ll 1$, during the inflationary epoch. In very good numerical precision, the minimal model essentially provides a realization of natural inflation \\begin{eqnarray} V(\\phi) \\approx \\Lambda^4 \\left[ 1-\\cos \\left( \\frac{\\phi}{\\feff} \\right) \\right] \\, , \\end{eqnarray} with $\\Lambda^4 = 9R^{-4}/(2\\pi)^6$ and $\\feff = (2\\pi\\gfd R)^{-1}$. For $10 \\lesssim R M_{\\rm Pl} \\lesssim 100$ and $1 \\lesssim \\feff/\\mpl \\lesssim 100$, the model predicts the observable cosmological quantities \\begin{equation} 1.2 \\times 10^{-5} \\lesssim \\mathcal{P}_\\mathcal{R} \\lesssim 4.9 \\times 10^{-5} \\, , \\end{equation} \\begin{equation} 0.952 \\lesssim n_\\mathcal{R} \\lesssim 0.966 \\, , \\end{equation} \\begin{equation} 0.03 \\lesssim r \\lesssim 0.13 \\, . \\end{equation} The power spectrum of the curvature perturbation $\\mathcal{P}_\\mathcal{R}$ and the corresponding spectral index $n_\\mathcal{R}$ are in good agreement with the current observations. While $|f_\\mathrm{NL}|$ is always far smaller than 1 and no detectable non-Gaussianity is expected, very interestingly the predicted tensor-to-scalar ratio $r$ is quite close to sensitivity of the near future cosmological experiments. This would be the first test of our minimal cosmological model. The reheating temperature $T_\\mathrm{RH}$ is estimated to be high enough to successfully follow the standard hot big bang evolution. \\subsection*{Acknowledgements} We are grateful to Misao Sasaki and the Yukawa Institute for Theoretical Physics at Kyoto University where some part of this work was carried out during ``Scientific Program on Gravity and Cosmology'' (YITP-T-07-01) and ``KIAS-YITP Joint Workshop: String Phenomenology and Cosmology'' (YITP-T-07-10). JG thanks Daniel Chung, L. Sriramkumar and Ewan Stewart for helpful conversations, and is partly supported by the Korea Research Foundation Grant KRF-2007-357-C00014 funded by the Korean Government. SCP appreciates Yasunori Nomura for his comments on low energy constraints of $\\gfd$ and also thanks C. S. Lim for encouragement to publish this paper." }, "0801/0801.1429_arXiv.txt": { "abstract": "We perform a combined X-ray and strong lensing analysis of RX~J1347.5-1145, one of the most luminous galaxy clusters at X-ray wavelengths. { We show that evidence from strong lensing alone, based on published VLT and new HST data, strongly argues in favor of a complex structure.} The analysis takes into account arc positions, shapes and orientations and is done thoroughly in the image plane. The cluster inner regions are well fitted by a bimodal mass distribution, with a total projected mass of $M_{tot} = (9.9 \\pm 0.3)\\times 10^{14} M_\\odot/h$ within a radius of $360~\\mathrm{kpc}/h$ ($1.5'$). Such a complex structure could be a signature of a recent major merger as further supported by X-ray data. A temperature map of the cluster, based on deep Chandra observations, reveals a hot front located between the first main component and an X-ray emitting South Eastern sub-clump. The map also unveils a filament of cold gas in the innermost regions of the cluster, most probably a cooling wake caused by the motion of the cD inside the cool core region. A merger scenario in the plane of the sky between two dark matter sub-clumps is consistent with both our lensing and X-ray analyses, and can explain previous discrepancies with mass estimates based on the virial theorem. ", "introduction": "\\label{intro} Accurate determination of total mass of galaxy clusters is important to understand properties and evolution of these systems, as well as for many cosmological applications. Gravitational lensing, through multiple image systems (strong lensing) as well as from distortions of background sources (weak lensing), provides a reliable method to determine the cluster mass, which is independent of the equilibrium properties of the cluster \\citep{mel99}. The lensing mass determination can be compared to estimates based on measured X-ray surface brightness and temperature, which is instead based on the assumption of hydrostatic equilibrium \\citep{sar88} or to dynamical estimates, which rely on the assumption of virialized systems. Combining optical, X-ray and radio observations of galaxy clusters is a major tool to investigate their intrinsic properties. In particular, the comparison of lensing and X-ray studies can give fundamental insights on the dynamical state of the galaxy clusters (see for example \\citet{allen2002, def+al04}), on the validity of the equilibrium hypothesis and on their 3-dimensional structure \\citep{def+al05, ser+al06, ser2007}. RX J1347.5-1145 ($z=0.451$) is one of the most X-ray luminous and massive galaxy cluster known. This cluster has been the subject of numerous X-ray \\citep{schindler95,sch+al97,allen2002,gi+sc04}, optical (Sahu et al.1998) and Sunyaev-Zeldovich (SZE) effect studies \\citep{kom+al01,kit+al04}. Formerly believed to be a well relaxed cluster, with a good agreement between weak-lensing \\citep{fi+ty97, kli+al05}, strong lensing \\citep{sahu98} and X-ray mass estimates \\citep{sch+al97}, more recent investigations revealed a more complex dynamical structure. { In particular}, a region of enhanced emission in the South-Eastern quadrant was first detected by SZE effect observations \\citep{kom+al01} and later confirmed by X-ray observations that also measured an hotter temperature for the excess component \\citep{allen2002}. This feature has been interpreted as an indication of a recent merger event \\citep{allen2002,kit+al04}. {Furthermore}, a spectroscopic survey on the cluster members found a velocity dispersion of $910 \\pm 130\\ \\mathrm{km~s^{-1}}$, which is significantly smaller than that derived from weak lensing, $1500 \\pm 160\\ \\mathrm{km~s^{-1}}$ (Fisher \\& Tyson 1997), strong lensing, roughly 1300 $\\mathrm{km~s^{-1}}$ and X-ray analyses $1320 \\pm 100\\ \\mathrm{km~s^{-1}}$ \\citep[see their table~4]{co+kn02}. A major merger in the plane of the sky was proposed as a likely scenario to reconcile all measurements \\citep{co+kn02}. In this paper, we further investigate the merger hypothesis by performing a combined strong lensing and X-ray analysis of archive data. We perform a strong lensing investigation based on a family of multiple arc candidates first proposed in Brada{\\v c} et al. (2005) using deep VLT observations. Differently from previous studies, we take care of performing the statistical analysis in the lens plane, which is a more reliable approach than the source-plane investigation when working with only one multiple image system. We further refine our analysis by taking into account not only the image positions, but also the shape and orientation of the arcs. In addition, we exploit {\\it Chandra} observations to gain additional insights into the dynamical status of the cluster, through spectral and morphological analyses of the X-ray halo, and to discriminate between different evolutionary scenarios. The paper is organised as follows. Section~\\ref{sec:data} discusses the strong lensing image candidates selection from archive VLT and { HST} data. Section~\\ref{pdsa} describes the statistical method used in the lensing analysis whereas Sec.~\\ref{Sec:xray} is devoted to the X-ray data analysis. Section~\\ref{sec:dis} discusses the merger hypothesis. Summary and conclusions are presented in Sect.~\\ref{sec:summary}. Throughout this paper we use a flat model of universe with a cosmological constant with $\\Omega_m=0.3$ and $H_0=70\\ \\mathrm{km\\ s}^{-1} \\mathrm{Mpc}^{-1}$. This implies a linear scale of $5.77\\ {\\rm kpc}/\\arcsec$ at the cluster redshift. ", "conclusions": "\\label{sec:summary} We have analysed both the lensing and X-ray properties of RX~J1347.5-11.45, one of the most luminous and massive X-ray cluster known. Based on the analysis of an arc family, photometrically selected, we have estimated the total cluster mass distribution, within a radius of $\\sim 500\\ \\rm{kpc}$ from the cluster centre. We performed a $\\chi^2$ analysis in the lens plane and scrupulously modelled the arc configuration. A model with two smooth dark matter components of similar mass accurately reproduces the observations and yields a mass estimate in agreement with previous strong and weak lensing and X-ray studies. Our strong lensing analysis suggests a major merger between two sub-clumps of similar mass located within the central $300~\\mathrm{kpc}$. X-ray observations further strengthen our view of a complex structure of the inner regions, revealing a hot front in the South-Eastern area, most probably a remnant of the occurred merger. This merging framework, which arises naturally from our strong lensing model, can also reconcile the observed discrepancy between dynamical mass estimates and X-ray, lensing and SZE ones, which instead give consistent results. Agreement between X-ray and lensing mass estimates further indicates that the gas might have had time to virialize. Spectroscopic measurements additionally suggest that the merger is taking place in the plane of the sky. Whereas the presence of a merger is confirmed on several grounds, its properties are though still unclear. The detailed features of our model are strictly related to the selection of the members of the multiple image system, in particular to the inclusion of the arc AC. Despite all candidates should be confirmed spectroscopically, a photometric analysis suggests that A4 and A5 almost undoubtedly belong to the same system. Remarkably, based on the shapes and orientations of A4 and A5 alone, without any constraint on AC, we can exclude the presence of a single mass component. A spectroscopic confirmation of further arc candidates is the required step to confirm and accurately define the merging scenario and to provide a final, more accurate description of the dynamical state of RX J1347.5-1145." }, "0801/0801.3276_arXiv.txt": { "abstract": "When constraining the primordial non-Gaussianity parameter $f_{\\rm NL}$ with cosmic microwave background anisotropy maps, the bias resulting from the covariance between primordial non-Gaussianity and secondary non-Gaussianities to the estimator of $f_{\\rm NL}$ is generally assumed to be negligible. We show that this assumption may not hold when attempting to measure the primordial non-Gaussianity out to angular scales below a few tens arcminutes with an experiment like Planck, especially if the primordial non-Gaussianity parameter is around the minimum detectability level with $f_{\\rm NL}$ between 5 and 10. In future, it will be necessary to jointly estimate the combined primordial and secondary contributions to the CMB bispectrum and establish $f_{\\rm NL}$ by properly accounting for the confusion from secondary non-Gaussianities. ", "introduction": " ", "conclusions": "" }, "0801/0801.1645.txt": { "abstract": " ", "introduction": "Models with warped extra dimensions~\\cite{Randall:1999ee} have arisen in the last few years as strong candidates for a natural theory of electroweak symmetry breaking (EWSB). The original solution to the hierarchy problem has been supplemented by the addition of a natural flavor structure~\\cite{flavor}, a custodial symmetry to protect the $T$ parameter~\\cite{Agashe:2003zs} and the $Z \\bar{b}_L b_L$ coupling~\\cite{Agashe:2006at}, and the realization of the Higgs as the pseudo Goldstone Boson of a broken global symmetry~\\cite{Contino:2003ve}--\\cite{Agashe:2005dk}. These developments have produced calculable models that successfully address most of the mysteries related to the electroweak (EW) scale. On the other hand, dark matter (DM), that is often considered one of the strongest --albeit indirect-- hints of physics beyond the Standard Model (SM), has so far lacked a generic implementation in models with warped extra dimensions. The main reason is the inherent asymmetry in warped backgrounds, that do not posses a natural KK parity as the one present in Universal Extra Dimensions (UED)~\\cite{Appelquist:2000nn}. In this article we explore a generic procedure to introduce an exact discrete exchange symmetry that results in new stable states, without introducing new parameters. This is done via a doubling of part of the field content. The exchange symmetry we advocate has been first introduced in \\cite{Panico:2006em} (where it was dubbed ``mirror symmetry'') and \\cite{Regis:2006hc} to alleviate the fine--tuning and to get a viable DM candidate in Gauge-Higgs Unification (GHU) models in flat space. As already anticipated in \\cite{Panico:2006em}, it can be extended straightforwardly to warped models. Given a bulk field, $\\phi$, satisfying certain boundary conditions (b.c.), the procedure consists of replacing $\\phi$ by a pair of fields, $\\phi_{1}$ and $\\phi_{2}$, and imposing the symmetry $\\phi_1 \\leftrightarrow \\phi_2$.~\\footnote{A similar discrete symmetry has recently been used in \\cite{Bai:2008cf} to get a DM candidate in the context of little Higgs theories.} The even linear combination $\\phi_+ \\equiv (\\phi_1 + \\phi_2)/\\sqrt{2}$ is identified with the original field (in particular, it inherits the b.c. obeyed by $\\phi$, as well as its couplings). The couplings of the orthogonal combination, $\\phi_- \\equiv (\\phi_1 - \\phi_2)/\\sqrt{2}$ are determined by those of $\\phi_{+}$. Under the exchange symmetry one has $\\phi_{\\pm}\\rightarrow \\pm \\phi_{\\pm}$, so that one can assign a multiplicative charge $+1$ to $\\phi_+$ and $-1$ to $\\phi_-$. Provided the discrete exchange symmetry is an exact symmetry at the quantum level, the lightest Kaluza--Klein (KK) resonance among all $\\Z_2$-odd states in the model is absolutely stable, and will be referred to as the LOP (Lightest Odd Particle). We argue that the above symmetry is indeed exact, by showing that possible 5D Chern--Simons (CS) terms, in general needed to restore gauge invariance in 5D theories, do not violate it~\\cite{Hill:2007zv}. The choice of which fields to double must be guided by phenomenological considerations. For example, DM direct and indirect searches impose stringent constraints on the possible couplings of the DM candidate to SM fields. It is therefore natural to look for charge and color neutral fields that can lead to viable DM candidates. In fact, the models we will consider always contain $U(1)$ factors, and it will be natural for the DM candidate to be a $U(1)$ massive gauge field, $X_{-}$. This is similar to the 5D UED case and the GHU model of \\cite{Regis:2006hc} in flat space, in which the DM can be identified with the first KK mode of the hypercharge gauge field. If the LOP were the only $\\Z_2$-odd particle, it would couple to SM fields only via non-renormalizable interactions. For the typical scales involved, its annihilation rate would then be extremely small, and the resulting thermal relic density unacceptably large. Hence, it is necessary in general to apply the previously described ``doubling'' construction to additional fields, making sure that $X_-$ remains the lightest among the $\\Z_{2}$-odd particles. We can choose b.c. for the LOP so that it has a mass about one order of magnitude smaller than the mass of the first $\\Z_2$-even gauge resonances. Thus, our DM candidate has typically sub--TeV masses, which, as we will see, can range from $\\sim 100$ GeV in Higgsless models up to $\\sim 700$ GeV in Randall--Sundrum (RS) models. Together with its mass, the coupling of the LOP to SM matter is the other crucial parameter governing its relic density. In all the cases we will consider, the $U(1)$ symmetry is related to the SM $U(1)_Y$ symmetry, with a coupling constant of electroweak size, and the $\\Z_{2}$-odd fields will consist of $X_-$ and a subset of fermion fields, typically associated with the top or bottom quarks. It is clear that our DM candidate is a weakly interacting massive particle (WIMP) with approximately the correct mass and couplings to give rise to the observed DM relic density in the universe. The construction outlined above turns out to be particularly natural in the specific class of GHU scenarios in warped space.\\footnote{It is known that such models can also be seen, thanks to the AdS/CFT dictionary \\cite{Ads-cft}, as strongly coupled 4D composite Higgs models. However, since calculability requires the 5D picture, we will mostly use the 5D language, adopting only occasionally the 4D dual language. } In this case, the additional set of fields introduced by our procedure is not only minimal, but is also singled out by the role the doubled sector plays in generating the EW scale dynamically. Although it is not easy to use the discrete symmetry to protect EW observables from large tree-level corrections, as in R-parity preserving supersymmetric scenarios, Little Higgs theories with T-parity, or UED with KK parity, we will see that the physics of EWSB in the GHU model with the discrete symmetry naturally leads to a region in parameter space where the constraints due to precision measurements are relaxed. Furthermore, this same region of parameter space, in which coannihilation effects can be relevant, predicts a dark matter abundance in accord with observation. Thus, the DM sector is tightly connected to the physics of EWSB, and plays an indirect role in leading to agreement with precision constraints.\\footnote{See \\cite{Hambye:2007vf} for a model in which EWSB and DM are related. In this reference, however, the hierarchy problem is not addressed.} Also, the model is rather predictive, with several fermionic resonances nearly degenerate with the DM particle that should lead to an exciting collider phenomenology. We also consider some non-perturbative effects --the formation of bound states and the effect of QCD Coulomb-like forces-- which might invalidate the usual perturbative computation of the relic density. We argue that the problem of the formation of bound states does not occur and that the effect of Coulomb--like forces on the perturbative cross-sections is negligible. We also discuss in some detail the degree of\u00cafine-tuning involved in the EWSB pattern, and compare to supersymmetric extensions of the standard model. It turns out that in the GHU framework (with or without DM) the fine-tuning is somewhat worse than expected by naive considerations. We point out, however, that here the fine-tuning seems to be associated with accommodating a top mass of order the EW scale, rather than with an intrinsic tension in the Higgs sector. When restricted to the region of parameter space with a fixed top mass, the low-energy properties of the model turn out to be essentially insensitive to the microscopic parameters of the model, and therefore to the detailed properties of the new physics, which is a very interesting feature. Although we find the application of our construction to the GHU framework particularly appealing, we stress that it is of more general applicability. We also briefly discuss the implementation of the discrete exchange symmetry in two other models with warped extra dimensions: the simplest RS model with the SM fermions and gauge bosons in the bulk, and a Higgsless model. Hence, our construction leads to viable DM candidates in a variety of scenarios, without the introduction of new parameters. The organization of the paper is as follows. In section~\\ref{sect:Z2} we review the generic properties of the exchange $\\Z_2$ symmetry that gives rise to a dark matter candidate. In section~\\ref{sec:GHU} we discuss in detail its implementation in a model of GHU. In particular, we study the interplay between EWSB, the EW constraints and the DM relic abundance, the (ir)relevance of the above mentioned non-perturbative effects, and discuss the fine-tuning. In section~\\ref{sec:othermodels} we describe the implementation of the discrete exchange symmetry in other models. Section~\\ref{sec:phenomenology} is devoted to particular details of DM collider phenomenology and DM direct detection in our construction. We comment on the issue of anomalies in section~\\ref{sect:anomalies}, and conclude in section~\\ref{sect:conclusions}. We relegate some technical details to the Appendices. %----------------------------------------------------------------------- ", "conclusions": "} In this paper we have proposed a generic construction that allows to endow given models with a DM candidate. This is achieved through an extension in which the model acquires a $\\Z_{2}$ exchange symmetry. The lightest $\\Z_{2}$-odd particle is then absolutely stable. We have considered several models with warped extra dimensions. Although these scenarios are well-motivated extensions of the SM, explaining both the Planck-weak scale hierarchy as well as the flavor structure of the SM, they do not contain, generically, stable particles that can account for the observed DM component of the universe. We have shown that our mechanism can easily solve this deficiency. We have payed special attention to a class of warped scenarios that is particularly appealing: GHU/composite Higgs scenarios. In this case, the $\\Z_{2}$ structure responsible for the stability of the DM candidate is tightly connected to the physics that leads to the dynamical breaking of the EW symmetry. As in supersymmetry, the dark matter mass and couplings are intimately connected to the EW scale. In fact, not only does our construction not introduce new parameters, but there is a further sense in which it can be considered minimal. As was emphasized in the main text, the physics of EWSB in such models crucially depends on certain fields that have two properties: they are fermionic fields without zero-modes and, through their strong connection to the top sector, they give a significant --in fact, crucial-- contribution to the Higgs potential. This is precisely the sector that gives rise to the $\\Z_{2}$-odd particles that can lead to a realistic dark matter candidate. As a result, the $\\Z_{2}$-odd sector, through its contribution to the Higgs potential, plays a key role in the realization of an EWSB vacuum with the desired physical properties. As explained in detail in the main text, the effects due to the $\\Z_{2}$-odd sector go in the direction of alleviating the constraints imposed by low-energy measurements. The picture that emerges is rather compelling: the observed DM relic abundance and the very precise measurements of the EW observables at LEP and SLC point to a \\textit{common} region in parameter space. This region is characterized by several fermionic resonances nearly degenerate with the DM candidate (a spin-1 particle), which in turn is predicted to have a mass in the $300-500~{\\rm GeV}$ range. We have also pointed out that although the GHU scenarios (with or without DM) present a sensitivity of order one percent or worse with respect to microscopic parameters (larger than what naive considerations would indicate), this sensitivity disappears once the top mass measurement is imposed. Thus, the moderate sensitivity of order a percent, that is present in virtually every extension of the SM that has any relation to the physics of EWSB acquires a new twist: it indicates an intrinsic difficulty in accommodating a heavy top. By contrast, once the top mass is fixed to a value of order the EW scale, the Higgs mass easily satisfies the LEP bound, while typically lying below $170~{\\rm GeV}$ or so. Regarding the phenomenological implications of our construction, the low scale predicted for the new $\\Z_2-$odd sector and, in particular, new vector-like quarks, guarantees large production cross-sections. Due to the particular features of the NLOP spectrum, which is very degenerate with the LOP, the most common signature will be jets plus missing energy accompanied in some cases by the semileptonic decays of a pair of off-shell $W^\\ast$. These are challenging signatures that will require dedicated analysis. The large production cross-sections and the information coming from the easier channels in the $\\Z_2-$even sector should however be sufficient to guarantee a discovery of the DM sector at colliders. We have also seen that DM direct detection is not very promising in present or near future experiments, due to the fact that couplings to light valence quarks are suppressed by intergenerational mixing, leading to very small cross-sections. \\vspace{5mm} \\noindent % %-----------------------------------------------------------------------" }, "0801/0801.4456_arXiv.txt": { "abstract": "{Hyper-velocity stars move so fast that only a supermassive black hole (SMBH) seems to be capable to accelerate them. Hence the Galactic centre (GC) is their only suggested place of origin. Edelmann et al.~(2005) found the early B-type star HE\\,0437$-$5439 to be too short-lived to have reached its current position in the Galactic halo if ejected from the GC, except if being a blue straggler star. Its proximity to the Large Magellanic Cloud (LMC) suggested an origin from this galaxy.} {The chemical signatures of stars at the GC are significantly different from those in the LMC. Hence, an accurate measurement of the abundance pattern of HE\\,0437$-$5439 will yield a new tight constraint on the place of birth of this hyper-velocity star.} {High-resolution spectra obtained with UVES on the VLT are analysed using state-of-the-art non-LTE modelling techniques.} {We measured abundances of individual elements to very high accuracy in HE\\,0437$-$5439 as well as in two reference stars, from the LMC and the solar neighbourhood, respectively. The abundance pattern is not consistent at all with that observed in stars near the GC, ruling our an origin from the GC. However, there is a high degree of consistency with the LMC abundance pattern. Our abundance results cannot rule out an origin in the outskirts of the Galactic disk. However, we find the life time of HE\\,0437$-$5439 to be more than~three times shorter than the time of flight to the edge of the disk, rendering a Galactic origin unlikely.} {Only one SMBH is known to be present in Galaxy and none in the LMC. Hence the exclusion of an GC origin challenges the SMBH paradigm. We conclude that there must be other mechanism(s) to accelerate stars to hyper-velocity speed than the SMBH. We draw attention to dynamical ejection from dense massive clusters, that has recently been proposed by Gvaramadze et al.~(2008).} ", "introduction": "Stars moving faster than the Galactic escape velocity were~first predicted to exist by \\citet{hills88}. The first such hyper-velocity stars~(HVSs) were discovered serendipitously only recently \\citep[][E05: \\object{HE\\,0437$-$5439} at 723\\,km\\,s$^{-1}$ heliocentric radial velocity]{brown05,hirsch05,edelmann05}. A systematic search for such objects has resulted in the discovery of seven additional HVS up to now \\citep[see][]{brown07}. \\cite{hills88} predicted that the tidal disruption of a binary by a supermassive black hole (SMBH) could lead to the ejection of stars with velocities exceeding the escape velocity of our Galaxy. The Galactic centre (GC) is the suspected place of origin of the HVSs as it hosts a SMBH. Both, the moderate rotational velocity ($v\\sin i$\\,$=$\\,54\\,\\kms) and the chemical composition (consistent with solar abundances within a factor of a few from a LTE analysis of a spectrum with S/N\\,$\\approx$\\,5) were considered evidence for a main sequence nature of HE\\,0437$-$5439 (E05). This put the star at a distance of $\\approx$60\\,kpc. Numerical kinematical experiments were carried out to trace the trajectory of HE\\,0437$-$5439 from the GC to its present location in the Galactic halo. However, its travel time (100\\,Myrs) was found to be much longer than its main sequence lifetime ($\\approx$\\,25--35\\,Myrs), rendering a GC origin unlikely, though possible if HE\\,0437$-$5439 would be a blue straggler star. Alternatively the star could have originated from the Large Magellanic Cloud (LMC) as it is much closer to this galaxy (18\\,kpc) than to the~GC. An accurate determination of its elemental abundance pattern should allow to distinguish between an origin in the LMC or in the GC because their chemical compositions differ distinctively. HE\\,0437$-$5439 is by far the brightest HVS known \\citep[$V$\\,$=$\\,16.36,][]{bonanos08} and therefore suited best for a detailed analysis. Hence we obtained high-resolution, high S/N spectra of HE\\,0437$-$5439 at the ESO VLT with UVES and performed a quantitative spectral analysis using state-of-the-art non-LTE modelling techniques. ", "conclusions": "We have performed a quantitative spectral analysis of the HVS HE\\,0437$-$3954 and reference objects in the LMC and the solar neighbourhood. Use of state-of-the-art non-LTE modelling techniques allowed precise elemental abundances to be obtained. From the results of the spectroscopic analysis we can rule out a GC origin for HE\\,0437$-$3954 because its abundance pattern indicates neither high metallicity nor $\\alpha$-enrichment. On the other hand a high degree of consistency with the LMC pattern is found when accounting for an abundance scatter within the LMC\\footnote{More comprehensive investigations than presently available, using improved analysis techniques on larger samples of young stars covering the entire region of the LMC are needed to draw definite conclusions on the degree of chemical (in)homogeneity in that galaxy.}. The observed abundance pattern could also be consistent with an origin from the outskirts of the Galactic disk, in the case HE\\,0437$-$3954 being a blue straggler. However, a comparison of the time-of-flight with the life time of the blue straggler practically rules out this scenario. Hence, HE\\,0437$-$3954 is unlikely of Galactic origin. \\citet{edelmann05} suggested that HE\\,0437$-$3954 was ejected from the LMC. Our spectroscopic analysis lends strong support to this scenario. The abundance pattern is consistent with that of our LMC reference star. An ejection velocity of about 1000\\,{\\kms} is required for the star to have reached its position from the LMC. Accurate proper motion measurements are needed to finally confirm the LMC origin. Because no SMBH is known to exist in the LMC, the SMBH slingshot scenario or other ejection models \\citep[e.g.][]{baumgardt06} invoking a SMBH are ruled out. Hence there must be an additional physical mechanism capable of accelerating a massive star to a space velocity of 1000\\,{\\kms}. Two such alternative scenarios have been proposed. {\\sc i}) A close encounter of a binary with an intermediate-mass black hole (IMBH) more massive than 10$^3$\\,$M_{\\sun}$ has been proposed as a viable ejection mechanism by \\citet{GPZ07}, who suggest the populous dense clusters \\object{NGC\\,2004} or \\object{NGC\\,2100}\\, in the LMC could harbour such an~IMBH. {\\sc ii}) Dynamical ejection by interaction of massive binaries in the cores of dense clusters is plausible \\citep{GGPZ08}. This process is more likely to occur in the LMC than in the Galaxy because the LMC hosts a significant number of sufficiently massive and dense clusters. From the list of \\citet{mackay03}, eight young clusters can be identified to be of the right age to qualify as candidate place of birth of HE\\,0437$-$3954. Primary candidates are NGC\\,2100 and NGC\\,2004, followed by \\object{NGC\\,1850} and \\object{NGC\\,1847}. In consequence, hyper-velocity stars are not such simple probes for the shape of the Galactic halo as suggested e.g. by \\citet{gnedin05}. The fundamental assumption of the exclusive origin of HVSs in the GC has to be dropped, as acceleration may occur in clusters throughout the Galactic disk as~well." }, "0801/0801.1363_arXiv.txt": { "abstract": "We develop a framework for studying collective three-flavor neutrino oscillations based on the density matrix formalism. We show how techniques proven useful for collective two-flavor neutrino oscillations such as corotating frames can be applied readily to three-flavor mixing. Applying two simple assumptions and the conservation of two ``lepton numbers'' we use this framework to demonstrate how the adiabatic/precession solution emerges. We illustrate with a numerical example how two stepwise spectral swaps appear naturally if the flavor evolution of the neutrino gas can be described by such a solution. For the special case where mu and tau flavor neutrinos are equally mixed and are produced with identical energy spectra and total numbers, we find that one of the spectral swaps in the three-flavor scenario agrees with that in the two-flavor scenario when appropriate mixing parameters are used. Using the corotating frame technique we show how the adiabatic/precession solution can obtain even in the presence of a dominant ordinary matter background. With this solution we can explain why neutrino spectral swapping can be sensitive to deviations from maximal 23-mixing when the ``mu-tau'' matter term is significant. ", "introduction": "It has long been recognized that, in addition to the conventional Mikheyev-Smirnov-Wolfenstein (MSW) effect \\cite{Wolfenstein:1977ue,Wolfenstein:1979ni,Mikheyev:1985aa}, neutrino self-coupling can be important for neutrino flavor evolution when neutrino number densities are large \\cite{Fuller:1987aa,Notzold:1988kx,Pantaleone:1992xh,Sigl:1992fn,% Fuller:1992aa,Qian:1993dg,Samuel:1993uw,Kostelecky:1994dt,Pastor:2001iu,% Pastor:2002we,Balantekin:2004ug,Fuller:2005ae}. Recently two-flavor neutrino oscillations in the core-collapse supernova environment have been intensively investigated \\cite{Duan:2005cp,Duan:2006an,Duan:2006jv,Hannestad:2006nj,% Raffelt:2007yz,EstebanPretel:2007ec,Duan:2007mv,Duan:2007bt,Fogli:2007bk}. These studies show that supernova neutrinos can indeed experience collective flavor evolution because of neutrino self-coupling, even when the neutrino self-coupling is subdominant compared to the MSW potential \\cite{Duan:2005cp}. An important result is that collective two-flavor neutrino oscillations can exhibit ``stepwise spectral swaps'' or ``spectral splits'' in the final neutrino energy spectra when the neutrino number density slowly decreases from a high value, where neutrinos experience synchronized oscillations \\cite{Pastor:2001iu}, towards zero (see, e.g., Ref.~\\cite{Duan:2006jv}). When this occurs, $\\nu_e$'s appear to swap their energy spectra with $\\nu_x$'s at energies below or above (depending on the neutrino mass hierarchy) a transition energy $\\Ec$ (where the superscript ``s'' can stand either for ``swapping point'' or ``splitting point''). Here $\\nu_x$ is some linear combination of $\\nu_\\mu$ and $\\nu_\\tau$. The phenomenon of spectral swapping is present in both the ``single-angle approximation'', where flavor evolution along various neutrino trajectories is assumed to be the same as that along a representative trajectory (e.g., the radial trajectory), and the ``multi-angle approximation'', where flavor evolution along different trajectories is independently followed \\cite{Duan:2006an}. For the inverted neutrino mass hierarchy, it also has been found that stepwise neutrino spectral swapping is essentially independent of the $2\\times2$ effective vacuum mixing angle when this angle is small (see, e.g., Ref.~\\cite{Duan:2007bt}). Collective two-flavor neutrino oscillations are best understood with the help of neutrino flavor polarization vectors \\cite{Sigl:1992fn} or neutrino flavor isospins \\cite{Duan:2005cp}. Using the spin analogy, one can represent the flavor content of a neutrino mode by a spin vector in flavor isospace. In this analogy, the matter effect is described as a spin-field coupling, and the neutrino self-coupling is described as spin-spin coupling. The phenomenon of spectral swapping is a result of collective precession of all neutrino flavor isospins with a common angular velocity $\\wpr$ at any given neutrino number density \\cite{Duan:2006an,Duan:2007mv}. This collective precession of neutrino flavor isospins is described the two-flavor adiabatic/precession solution \\cite{Raffelt:2007cb,Raffelt:2007xt}. In this solution, all neutrino flavor isospins stay aligned or antialigned with a total effective field in a reference frame that rotates with angular velocity $\\wpr$. Numerical simulations have shown that, in the supernova environment, neutrinos can first experience collective MSW-like flavor transformation (in which the MSW effect is enhanced by neutrino self-coupling) and then subsequently the adiabatic/precession solution \\cite{Duan:2007fw}. In the real world, however, there are three active neutrino flavors. Some limited progress has been made on understanding collective three-flavor neutrino oscillations. The first fully-coupled simulation of three-flavor neutrino oscillations in the supernova environment showed a spectral swapping phenomenon similar to the two-flavor scenario except possibly with two swaps \\cite{Duan:2007sh}. Another simplified numerical study with a single neutrino energy bin \\cite{EstebanPretel:2007yq} reveals that collective neutrino oscillations can be sensitive to deviations from maximal 23-mixing when there is a dominant ``mu-tau'' term arising from a high order contribution from virtual $\\mu$'s and $\\tau$'s (e.g., Refs.~\\cite{Fuller:1987aa,Botella:1987aa,Roulet:1995qb}). Ref.~\\cite{Dasgupta:2007ws} has extended the neutrino flavor polarization vector notation to the three-flavor scenario and discussed the collective three-flavor oscillations as ``factorization'' of two two-flavor oscillations. In this paper we develop a framework for studying collective three-flavor neutrino oscillations based on the density matrix formalism. Using this framework we find a generalized three-flavor adiabatic/precession solution and show how stepwise spectral swapping can appear as a natural result of such a solution. The rest of this paper is organized as follows. In Sec.~\\ref{sec:eom} we develop a framework centered around a $3\\times3$ reduced neutrino flavor density matrix. This density matrix is equivalent to an 8-component neutrino flavor vector, a generalization of neutrino flavor isospin and similar to the Bloch vector used in Ref.~\\cite{Dasgupta:2007ws}. We show how techniques important in studying collective two-flavor oscillations such as corotating frames can be applied in the framework. In Sec.~\\ref{sec:sol} we demonstrate how the three-flavor adiabatic/precession solution can be found using two simple assumptions and the conservation of two ``lepton numbers''. We also illustrate with a numerical example how two spectral swaps can form from the adiabatic/precession solution when the total neutrino number density vanishes. In Sec.~\\ref{sec:matt} we employ the corotating frame technique and show that the adiabatic/precession solution obtains even in the presence of a dominant matter background. In particular, we show that, in the presence of a large mu-tau term, neutrino spectral swapping becomes sensitive to deviations from maximal 23-mixing. In Sec.~\\ref{sec:conclusions} we give our conclusions. ", "conclusions": "} We have developed a framework for studying collective three-flavor neutrino oscillations. Important techniques in studying collective two-flavor oscillations such as corotating frames can be applied readily to three-flavor scenarios in this framework. We have shown that the three-flavor adiabatic/precession solution obtains when both the precession ansatz and the adiabatic ansatz are satisfied. If the flavor evolution of a neutrino gas is described by the adiabatic/precession solution, the final neutrino energy spectra will exhibit the stepwise swapping phenomenon. We have shown that stepwise spectral swapping appears in a numerical example in which neutrinos are directly emitted from the neutrino sphere of a bare neutron star into vacuum. For this special example, because the neutrinos in $\\mu$ and $\\tau$ flavors are equally mixed and have identical energy spectra initially, the adiabatic/precession solutions for both the three-flavor and the two-flavor scenarios produce the same spectral swapping at the atmospheric neutrino mass-squared-difference scale. In more general cases, however, this may not be the case, and the full $3\\times3$ mixing framework should be employed. Strictly speaking, the adiabatic/precession solution obtains only when the $CP$-violating phase $\\delta=0$. This is because, if $\\delta\\neq0$, the unitary transformation matrix $U$ connecting the flavor states and vacuum mass eigenstates is not real, and \\begin{equation} U\\rho_\\bfp^*U^\\dagger\\neq(U\\rho_\\bfp U^\\dagger)^* \\text{ and } U\\bar\\rho_\\bfp^*U^\\dagger\\neq(U\\bar\\rho_\\bfp U^\\dagger)^*. \\end{equation} As a result, Eqs.~\\eqref{eq:eom-nu} and \\eqref{eq:eom-anu} become invalid in the vacuum mass basis, and all the following derivations are invalid. In practice, however, the adiabatic/precession solution may still be a good approximation even when $\\delta$ is large. This is because $\\theta_{13}\\simeq0$ and the transformation matrix $U$ is almost real. In fact, numerical simulations in Ref.~\\cite{Duan:2007sh} show that, at least for the parameters employed in those simulations, varying the $CP$ phase $\\delta$ has little effect on the final neutrino energy spectra except for changing the relative mixing of $\\mu$ and $\\tau$ neutrino flavors. The possible effects of $CP$ violation in stellar collapse have been discussed in a different context (see, e.g., Ref.~\\cite{Balantekin:2007es}). We also have demonstrated that the adiabatic/precession solution obtains even in the presence of a dominant matter background and a large mu-tau term. When the matter term is much larger than the vacuum mixing term, the presence of the ordinary matter only reshuffles the neutrino states in which the spectral swapping has the most dramatic manifestation. For the supernova environment, this means that the regime where collective neutrino oscillations occur is solely determined by the neutrino fluxes and does not depend sensitively on the matter density profile. This is in agreement with the previous analysis in two-flavor scenarios \\cite{Duan:2005cp}. The supernova neutrino signals observed on earth will depend of course on the matter profile in the supernova. In part, this is because the MSW effect will modify the neutrino energy spectra subsequent to the collective oscillations discussed here. This dependence of neutrino signals on the matter profile can provide important information on the conditions deep in the supernova envelope \\cite{Schirato:2002tg,Duan:2007sh,Lunardini:2007vn}." }, "0801/0801.1680_arXiv.txt": { "abstract": "{Rare types of variable star may give unique insight into short-lived stages of stellar evolution. The systematic monitoring of millions of stars and advanced light curve analysis techniques of microlensing surveys make them ideal for discovering also such rare variable stars. One example is the R Coronae Borealis (RCB) stars, a rare type of evolved carbon-rich supergiant.} {We have conducted a systematic search of the EROS-2 database for the Galactic catalogue Bulge and spiral arms to find Galactic RCB stars.} {The light curves of $\\sim$100 million stars, monitored for 6.7 years (from July 1996 to February 2003), have been analysed to search for the main signature of RCB stars, large and rapid drops in luminosity. Follow-up spectroscopy has been used to confirm the photometric candidates.} {We have discovered 14 new RCB stars, all in the direction of the Galactic Bulge, bringing the total number of confirmed Galactic RCB stars to about 51.} {After reddening correction, the colours and absolute magnitudes of at least 9 of the stars are similar to those of Magellanic RCB stars. This suggests that these stars are in fact located in the Galactic Bulge, making them the first RCB stars discovered in the Bulge. The localisation of the 5 remaining RCBs is more uncertain: 4 are either located behind the Bulge at an estimated maximum distance of 14 kpc or have an unusual thick circumstellar shell; the other is a DY Per RCB which may be located in the Bulge, even if it is fainter than the known Magellanic DY Per. From the small scale height found using the 9 new Bulge RCBs, $61